Hypochlorous acid

Jump to navigation Jump to search

Template:Chembox new

Hypochlorous acid (IUPAC name chloric(I) acid) is a weak acid with the chemical formula HClO. It forms when chlorine dissolves in water. It cannot be isolated in pure form due to rapid equilibration with its precursor (see below). HClO is used as a bleach, an oxidizer, a deodorant, and a disinfectant.

Formation

Addition of chlorine to water gives both chloric(I) acid and hydrochloric acid[1] (HCl):

Cl2 + H2O → HOCl + HCl

Uses

In organic synthesis, HOCl converts alkenes to chlorohydrins.[2]

In biology, hypochlorous acid is generated in activated neutrophils by myeloperoxidase mediated peroxidation of chloride ions, and contributes to the destruction of bacteria.[3][4][5]

Hypochlorous acid is the active sanitizer in hypochlorite based swimming pool products.

Chemical reactions

In aqueous solution, hypochlorous acid partially dissociates into the anion hypochlorite ClO-:

HOCl<math>\overrightarrow{\leftarrow}</math> OCl- + H+

Salts of hypochlorous acid are also called hypochlorites. One of the best known hypochlorites is NaOCl, the active ingredient in bleach.

In the presence of sunlight, hypochlorous acid decomposes into hydrochloric acid and oxygen, so this reaction is sometimes seen as:

2Cl2 + 2H2O → 4HCl + O2

HOCl is considered to be a stronger oxidant than chlorine.

HOCl reacts with HCl to form chlorine gas:

HOCl + HCl → H2O + Cl2

Reactivity of HOCl with biomolecules

Hypochlorous acid reacts with a wide variety of biomolecules including DNA, RNA,[5][6][7][8] fatty acid groups, cholesterol[9][10][11][12][13][14][15][16] and proteins.[1][17][12][18][19][20][21]

Reaction with protein sulfhydral groups

Knox et al.[19] first noted that HOCl was a sulfhydral inhibitor that in sufficient quantity could completely inactivate proteins containing sulfhydral groups. This is because, HOCl oxidises sulfhydral groups leading to the formation of disulfide bonds[22] that can result in crosslinking of proteins. The HOCl mechanism of sulfhydral oxidation is similar to that of chloramine, and may only be bacteriostatic, because once the residual chlorine is dissipated, some sulfhydral function can be restored.[18] One sulfhydral containing amino acid can scavenge up to four molecules of HOCl.[21] Consistent with this, it has been proposed that sulfhydral groups of sulfur containing amino acids can be oxidized a total of three times by three HOCl molecules, with the fourth reacting with the α-amino group. The first reaction yields sulfenic acid (R-SOH) then sulfinic acid (R-SO2H) and finally R-SO3H. Each of those intermediates can also condense with another sulfhydral group causing cross linking and aggregation of proteins. Sulfinic acid and R-SO3H derivatives are only produced at high molar excesses of HOCl, and disulfides are primarily formed at bacteriocidal levels.[8] Disulfide bonds can also be oxidized by HOCl to sulfinic acid.[22] Because the oxidation of sulfhydrals and disulfides evolves hydrochloric acid,[8] this process results in the depletion HOCl.

Reaction with protein amino groups

Hypochlorous acid reacts readily with amino acids that have amino group side chains, with the chlorine from HOCl displacing a hydrogen resulting in an organic chloramine.[23] Chlorinated amino acids rapidly decompose but protein chloramines are longer lived and retain some oxidative capacity.[21][4] Thomas et al.[4] concluded from their results that most organic chloramines decayed by internal rearrangement and that fewer available NH2 groups promoted attack on the peptide bond resulting in cleavage of the protein. McKenna and Davies[24] found that 10 mM or greater HOCl was necessary to fragment proteins in vivo. Consistent with these results it was later proposed that the chloramine undergoes a molecular rearrangement releasing HCl and ammonia to form an amide.[25] The amide group can further react with another amino group to form a Schiff base causing cross linking and aggregation of proteins.[12]

Reaction with DNA and Nucleotides

Hypochlourous acid reacts slowly with DNA and RNA as well as all nucleotides in vitro.[6][26] GMP is the most reactive because HOCl reacts with both the heterocyclic NH group and the amino group. Similarly TMP with only a heterocyclic NH group that is reactive with HOCl is the second most reactive. AMP and CMP which only have a slowly reactive amino group are less reactive with HOCl.[26] UMP has been reported to be reactive only at a very slow rate.[5][6] The heterocyclic NH groups are more reactive than amino groups and their secondary chloramines are able to donate the chlorine.[8] These reactions likely interfere with DNA base pairing and consistent with this, Prütz[26] has reported a decrease in viscosity of DNA exposed to HOCl similar to that seen with heat denaturation. The sugar moieties are unreactive and the DNA backbone is not broken.[26] NADH can react with chlorinated TMP and UMP as well as HOCl. This reaction can regenerate UMP and TMP and results in the 5-hydroxy derivative of NADH. The reaction with TMP or UMP is slowly reversible to regenerate HOCl. A second slower reaction that results in cleavage of the pyridine ring occurs when excess HOCl is present. NAD+ is inert to HOCl.[26][8]

Reaction with lipids

Hypochlorous acid reacts with unsaturated bonds in lipids, but not saturated bonds, and the OCl ion does not participate in this reaction. This reaction occurs by hydrolysis with addition of chlorine to one of the carbons and a hydroxyl to the other. The resulting compound is a chlorhydrin.[9] The polar chlorine disrupts lipid bilayers and could increase permeability.[10] When chlorhydrin formation occurs in lipid bilayers of red blood cells, increased permeability occurs. Disruption could occur if enough chlorhydrin is formed.[9][15] The addition of preformed chlorhydrins to red blood cells can affect permeability as well.[11] Cholesterol chlorhydrins have also been observed,[10][13] but do not greatly affect permeability, and it is believed that Cl2 is responsible for this reaction.[13]

Mode of disinfectant action

Escherichia coli exposed to hypochlorous acid lose viability in less than 100 ms due to inactivation of many vital systems.[1][27][28][29][30] Hypochlorous acid has a reported LD50 of 0.0104 ppm - 0.156 ppm[31] and 2.6 ppm caused 100% growth inhibition in 5 minutes.[24] However it should be noted that the concentration required for bactericidal activity is also highly dependent on bacterial concentration.[19]

Inhibition of glucose oxidation

In 1948, Knox et al.[19] proposed the idea that inhibition of glucose oxidation was a major factor in the bacteriocidal nature of chlorine solutions. He proposed that the active agent or agents diffused across the cytoplasmic membrane to inactivate key sulfhydral containing enzymes in the glycolytic pathway. This group was also the first to note that chlorine solutions (HOCl) inhibited sulfhydral enzymes. Later studies have shown that at bacteriocidal levels, the cytosol components do not react with HOCl.[32] In agreement with this, McFeters and Camper[33] found that aldolase, an enzyme that Knox et al.[19] proposes would be inactivated, was unaffected by HOCl in vivo. It has been further shown that loss of sulfhydrals does not correlate with inactivation.[18] That leaves the question what causes inhibition of glucose oxidation. The discovery that HOCl blocks induction of β-galactosidase by added lactose[34] led to a possible answer to this question. The uptake of radiolabeled substrates by both ATP hydrolysis and proton co-transport may be blocked by exposure to HOCl preceding loss of viability.[32] From this observation it proposed that HOCl blocks uptake of nutrients by inactivating transport proteins.[32][17][35][33] The question of loss of glucose oxidation has been further explored in terms of loss of respiration. Venkobachar et al.[36] found that succinic dehydrogenase was inhibited in vitro by HOCl and this led to the investigation of the possibility that disruption of electron transport could be the cause of bacterial inactivation. Albrich et al.[5] subsequently found that HOCl destroys cytochromes and iron-sulfur clusters and observed that oxygen uptake is abolished by HOCl and adenine nucleotides are lost. Also observed was, that irreversible oxidation of cytochromes paralleled the loss of respiratory activity. One way of addressing the loss of oxygen uptake was by studying the effects of HOCl on succinate dependent electron transport.[37] Rosen et al.[30] found that levels of reductable cytochromes in HOCl treated cells were normal, and these cells were unable to reduce them. Succinate dehydrogenase was also inhibited by HOCl, stopping the flow of electrons to oxygen. Later studies[28] revealed that Ubiquinol oxidase activity ceases first, and the still active cytochromes reduce the remaining quinone. The cytochromes then pass the electrons to oxygen, which explains why the cytochromes cannot be reoxidized as observed by Rosen et al.[30] However, this line of inquiry was ended when Albrich et al.[1] found that cellular inactivation precedes loss of respiration by using a flow mixing system that allowed evaluation of viability on much smaller time scales. This group found that cells capable of respiring could not divide after exposure to HOCl.

Depletion of adenine nucleotides

Having eliminated loss of respiration Albrich et al.[1] proposes that the cause of death may be due to metabolic dysfunction caused by depletion of adenine nucleotides. Barrette et al.[34] studied the loss of adenine nucleotides by studying the energy charge of HOCl exposed cells and found that cells exposed to HOCl were unable to step up their energy charge after addition of nutrients. The conclusion was that exposed cells have lost the ability to regulate their adenylate pool, based on the fact that metabolite uptake was only 45% deficient after exposure to HOCl and the observation that HOCl causes intracellular ATP hydrolysis. Also confirmed was; that at bacteriocidal levels of HOCl, cytosolic components are unaffected. So it was proposed that modification of some membrane bound protein results in extensive ATP hydrolysis, and this, coupled with the cells inability to remove AMP from the cytosol depresses metabolic function. One protein involved in loss of ability to regenerate ATP has been found to be ATP synthetase.[17] Much of this research on respiration reconfirms the observation that relevant bacteriocidal reactions take place at the cell membrane.[34][17][38]

Inhibition of DNA replication

Recently it has been proposed that bacterial inactivation by HOCl is the result of inhibition of DNA replication. When bacteria are exposed to HOCl there is a precipitous decline in DNA synthesis that precedes inhibition of protein synthesis, and closely parallels loss of viability.[24][39] During bacterial genome replication, the origin of replication (oriC in E. coli) binds to proteins that are associated with the cell membrane, and it was observed that HOCl treatment decreases the affinity of extracted membranes for oriC, and this decreased affinity also parallels loss of viability. A study by Rosen et al[40] compared the rate of HOCl inhibition of DNA replication of plasmids with different replication origins and found that certain plasmids exhibited a delay in the inhibition of replication when compared to plasmids containing oriC. Rosen’s group proposed that inactivation of membrane proteins involved in DNA replication are the mechanism of action of HOCl.

Safety

HOCl is a strong oxidizer and can form explosive mixtures.

External links

References

  1. 1.0 1.1 1.2 1.3 1.4 Fair, G. M., J. C. Morris, S. L. Chang, I. Weil, and R. P. Burden. 1948. The behavior of chlorine as a water disinfectant. J. Am. Water Works Assoc. 40:1051-1061.
  2. Unangst, P. C. "Hypochlorous Acid" in Encyclopedia of Reagents for Organic Synthesis (Ed: L. Paquette) 2004, J. Wiley & Sons, New York. DOI: 10.1002/047084289.
  3. Harrison, J. E., and J. Schultz. 1976. Studies on the chlorinating activity of myeloperoxidase. Journal of Biological Chemistry volume 251, pages1371-1374.
  4. 4.0 4.1 4.2 Thomas, E. L. 1979. Myeloperoxidase, hydrogen peroxide, chloride antimicrobial system: Nitrogen-chlorine derivatives of bacterial components in bactericidal action against Escherichia coli. Infect. Immun. 23:522-531.
  5. 5.0 5.1 5.2 5.3 Albrich, J. M., C. A. McCarthy, and J. K. Hurst. 1981. Biological reactivity of hypochlorous acid: Implications for microbicidal mechanisms of leukocyte myeloperoxidase. Proc. Natl. Acad. Sci. USA 78:210-214.
  6. 6.0 6.1 6.2 Dennis, W. H., Jr, V. P. Olivieri, and C. W. Krusé. 1979. The reaction of nucleotides with aqueous hypochlorous acid. Water Res. 13:357-362.
  7. Jacangelo, J. G., and V. P. Olivieri. 1984. Aspects of the mode of action of monochloramine. In R. L. Jolley, R. J. Bull, W. P. Davis, S. Katz, M. H. Roberts, Jr., and V. A. Jacobs (ed.), Water Chlorination, vol. 5. Lewis Publishers, Inc., Williamsburg.
  8. 8.0 8.1 8.2 8.3 8.4 Prütz, W. A. 1998. Interactions of hypochlorous acid with pyrimidine nucleotides, and secondary reactions of chlorinated pyrimidines with GSH, NADPH, and other substrates. Arch. Biochem. Biophys. 349:183-191.
  9. 9.0 9.1 9.2 Arnhold, J., O. M. Panasenko, J. Schiller, Y. A. Vladimirov, and K. Arnold. 1995. The action of hypochlorous acid on phosphatidylcholine liposomes in dependence on the content of double bonds. Stoichiometry and NMR analysis. Chem. Phys. Lipids 78:55-64.
  10. 10.0 10.1 10.2 Carr, A. C., J. V. D. Berg, and C. C. Winterbourn. 1996. Chlorination of cholesterol in cell membranes by hypochlorous acid. Arch. Biochem. Biophys. 332:63-69.
  11. 11.0 11.1 Domigan, N. M., M. C. M. Vissers, and C. C. Winterbourn. 1997. Modification of red cell membrane lipids by hypochlorous acid and haemolysis by preformed lipid chlorhydrins. Redox Rep. 3:263-271.
  12. 12.0 12.1 12.2 Hazell, L. J., J. V. D. Berg, and R. Stocker. 1994. Oxidation of low density lipoprotein by hypochlorite causes aggregation that is mediated by modification of lysine residues rather than lipid oxidation. Biochem. J. 302:297-304.
  13. 13.0 13.1 13.2 Hazen, S. L., F. F. Hsu, K. Duffin, and J. W. Heinicke. 1996. Molecular chlorine generated by the myeloperoxidase-hydrogen peroxide-chloride system of phagocytes converts low density lipoprotein cholesterol into a family of chlorinated sterols. J. Biol. Chem. 271:23080-23088.
  14. Vissers, M. C. M., A. C. Carr, and A. L. P. Chapman. 1998. Comparison of human red cell lysis by hypochlorous acid and hypobromous acids: insights into the mechanism. Biochem. J. 330:131-138.
  15. 15.0 15.1 Vissers, M. C. M., A. Stern, F. Kuypers, J. V. D. Berg, and C. C. Winterbourn. 1994. Membrane changes associated with lysis of red blood cells by hypochlorous acid. Free Rad. Biol. Med. 16:703-712.
  16. Winterbourne, C. C., J. V. D. Berg, E. Roitman, and F. A. Kuypers. 1992. Chlorhydrin formation from unsaturated fatty acids reacted with hypochlorous acid. Arch. Biochem. Biophys. 296:547-555.
  17. 17.0 17.1 17.2 17.3 Barrette, W. C., Jr., D. M. Hannum, W. D. Wheeler, and J. K. Hurst. 1989. General mechanism for the bacterial toxicity of hypochlorous acid: Abolition of ATP production. Biochemistry 28:9172-9178.
  18. 18.0 18.1 18.2 Jacangelo, J. G., V. P. Olivieri, and K. Kawata. 1987. Oxidation of sulfhydryl groups by monochloramine. Water Res. 21:1339-1344.
  19. 19.0 19.1 19.2 19.3 19.4 Knox, W. E., P. K. Stumpf, D. E. Green, and V. H. Auerbach. 1948. The inhibition of sulfhydral enzymes as the basis of the bactericidal action of chlorine. J. Bacteriol. 55:451-458.
  20. Vissers, M. C. M., and C. C. Winterbourne. 1991. Oxidative Damage to Fibronectin. Arch. Biochem. Biophys. 285:53-59.
  21. 21.0 21.1 21.2 Winterbourne, C. C. 1985. Comparative reactivities of various biological compounds with myeloperoxidase-hydrogen peroxide-chloride, and similarity to the oxidant to hypochlorite. Biochim. Biophys. Acta 840:204-210.
  22. 22.0 22.1 Pereira, W. E., Y. Hoyano, R. E. Summons, V. A. Bacon, and A. M. Duffield. 1973. Chlorination studies: II. The reaction of aqueous hypochlorous acid with a - amino acids and dipeptides. Biochim. Biophys. Acta 313:170-180.
  23. Dychdala, G. R. 1991. Chlorine and chlorine compounds, p. 131-151. In S. S. Block (ed.), Disinfection, Sterilization and Preservation. Lea & Febiger, Philadelphia.
  24. 24.0 24.1 24.2 McKenna, S. M., and K. J. A. Davies. 1988. The inhibition of bacterial growth by hypochlorous acid: possible role in the bactericidal activity of phagocytes. Biochem. J. 254:685-692.
  25. Hazen, S. L., A. d'Avignon, M. M. Anderson, F. F. Hsu, and J. W. Heinicke. 1998. Human neutrophils employ the myeloperoxidase-hydrogen peroxide-chloride system to oxidize α-amino acids to a family of reactive aldehydes. J. Biol. Chem. 273:4997-5005.
  26. 26.0 26.1 26.2 26.3 26.4 Prütz, W. A. 1996. Hypochlorous acid interactions with thiols, nucleotides, DNA and other biological substrates. Arch. Biochem. Biophys. 332:110-120.
  27. Rakita, R. M., B. R. Michel, and H. Rosen. 1990. Differential inactivation of Escherichia coli membrane dehydrogenases by a myeloperoxidase-mediated antimicrobial system. Biochemistry 29:1075-1080.
  28. 28.0 28.1 Rakita, R. M., B. R. Michel, and H. Rosen. 1989. Myeloperoxidase-mediated inhibition of microbial respiration: Damage to Escherichia coli ubiquinol oxidase. Biochemistry 28:3031-3036.
  29. Rosen, H., and S. J. Klebanoff. 1985. Oxidation of microbial iron-sulfur centers by the myeloperoxidase-H2O2-halide antimicrobial system. Infect. Immun. 47:613-618.
  30. 30.0 30.1 30.2 Rosen, H., R. M. Rakita, A. M. Waltersdorph, and S. J. Klebanoff. 1987. Myeloperoxidase-mediated damage to the succinate oxidase system of Escherichia coli. J. Biol. Chem. 242:15004-15010.
  31. Chesney, J. A., J. W. Eaton, and J. R. Mahoney, Jr. 1996. Bacterial glutathione: a sacrificial defense against chlorine compounds. J. Bacteriol. 178:2131-2135.
  32. 32.0 32.1 32.2
  33. 33.0 33.1 McFeters, G. A., and A. K. Camper. 1983. Enumeration of indicator bacteria exposed to chlorine. Adv. Appl. Microbiol. 29:177-193.
  34. 34.0 34.1 34.2 Barrette, W. C., Jr., J. M. Albrich, and J. K. Hurst. 1987. Hypochlorous acid-promoted loss of metabolic energy in Escherichia coli. Infect. Immun. 55:2518-2525.
  35. Camper, A. K., and G. A. McFeters. 1979. Chlorine injury and the enumeration of waterborne coliform bacteria. Appl. Environ. Microbiol. 37:633-641.
  36. Venkobachar, C., L. Iyengar, and A. V. S. P. Rao. 1975. Mechanism of disinfection. Water Res. 9:119-124.
  37. Hurst, J. L., W. C. Barrette, Jr., B. R. Michel, and H. Rosen. 1991. Hypochlorous acid and myeloperoxidase-catalyzed oxidation of iron-sulfur clusters in bacterial respiratory dehydrogenases. Eur. J. Biochem. 202:1275-1282.
  38. Rosen, H., and S. J. Klebanoff. 1982. Oxidation of Escherichia coli iron centers by myeloperoxidase-mediated microbicidal system. J. Biol. Chem. 257:13731-13735
  39. Rosen, H., J. Orman, R. M. Rakita, B. R. Michel, and D. R. VanDevanter. 1990. Loss of DNA-membrane interactions and cessation of DNA synthesis in myeloperoxidase-treated Escherichia coli. Proc. Natl. Acad. Sci. USA 87:10048-10052.
  40. Rosen, H., B. R. Michel, D. R. vanDevanter, and J. P. Hughes. 1998. Differential effects of myeloperoxidase-derived oxidants on Escherichia coli DNA replication. Infect. Immun. 66:2655-2659.

de:Hypochlorige Säure ko:하이포아염소산 it:Acido ipocloroso hu:Hipoklórossav nl:Waterstofhypochloriet sl:Hipoklorova kislina sr:Хипохлораста киселина uk:Гіпохлоритна кислота