Reactive_oxygen_species_production_in_marine_microalgae

Reactive oxygen species production in marine microalgae

Reactive oxygen species production in marine microalgae

Add article description


All living cells produce reactive oxygen species (ROS) as a byproduct of metabolism. ROS are reduced oxygen intermediates that include the superoxide radical (O2) and the hydroxyl radical (OH•), as well as the non-radical species hydrogen peroxide (H2O2). These ROS are important in the normal functioning of cells, playing a role in signal transduction[1][2] and the expression of transcription factors.[3][4] However, when present in excess, ROS can cause damage to proteins, lipids and DNA by reacting with these biomolecules to modify or destroy their intended function. As an example, the occurrence of ROS have been linked to the aging process in humans, as well as several other diseases including Alzheimer's, rheumatoid arthritis, Parkinson's, and some cancers.[5] Their potential for damage also makes reactive oxygen species useful in direct protection from invading pathogens,[6] as a defense response to physical injury,[7][8][9][10] and as a mechanism for stopping the spread of bacteria and viruses by inducing programmed cell death.[11]

Reactive oxygen species are present in low concentrations in seawater and are produced primarily through the photolysis of organic and inorganic matter.[12] However, the biological production of ROS, generated through algal photosynthesis and subsequently 'leaked' to the environment, can contribute significantly to concentrations in the water column.[13][14][15] Although there is very little information on the biological generation of ROS in marine surface waters, several species of marine phytoplankton have recently been shown to release significant amounts of ROS into the environment.[16][17] This ROS has the potential to harm nearby organisms,[18][19] and, in fact, has been implicated as the cause of massive fish, bacteria, and protist mortalities.[20][21][22]

Chemical background

In sea water, ROS can be generated through abiotic as well as biotic processes, among which are the radiolysis and photolysis of water molecules and cellular respiration. According to a model proposed by Fan[23] for the prediction of ROS in surface waters, the biochemistry mediated by phytoplankton may be just as important for the production of ROS as photochemistry. Biological ROS is often synthesized in mitochondrial membranes, as well as the endoplasmic reticulum of animals, plants, and some bacteria.[24][25] In addition, chloroplasts and the organelles peroxisomes and glyoxysomes are also sites for the generation of ROS.[24][25][26] The ROS most likely released to the environment are those produced at the cell surface as electrons get "leaked" from the respiratory chain and react with molecular oxygen, O2.[27] The products of this subsequent reduction of molecular oxygen are what are referred to as reactive oxygen species. Thus, the production of ROS is in direct proportion to the concentration of O2 in the system, with increases of O2 leading to higher production of ROS.[28] There are three main reactive oxygen species: the superoxide anion (O2), hydrogen peroxide (H2O2), and the hydroxyl radical (OH•). The superoxide anion is formed directly from the one-electron reduction of molecular oxygen.[29] Hydrogen peroxide is then formed from the disproportionation of the superoxide anion. This reaction occurs very quickly in seawater. Next, the reduction of hydrogen peroxide yields the hydroxyl radical, H2O2↔2OH•, which can then get reduced to the hydroxyl ion and water.[1] However, the presence of reactive oxygen species in marine systems is hard to detect and measure accurately, for a number of reasons. First, ROS concentrations are generally low (nanomoles) in seawater. Second, they may react with other hard to identify molecules that occur in low quantities, resulting in unknown products. Finally, they are (for the most part) transient intermediates, having lifetimes as little as microseconds.[30]

Superoxide production

According to Blough & Zepp,[30] superoxide is one of the hardest reactive oxygen species to quantify because it is present in low concentrations: 2×10−12 M in the open ocean and up to 2×10−10M in coastal areas. The main sources of biological superoxide in the ocean come from the reduction of oxygen at the cell surface and metabolites released into the water.[27][31] In marine systems, superoxide most often acts as a one-electron reductant, but it can also serve as an oxidant and may increase the normally slow oxidation rates of environmental compounds.[12][30] Superoxide is very unstable, with between 50 and 80% of its concentration of anions spontaneously disproportionating to hydrogen peroxide. At its peak, this reaction occurs with a rate constant on the order of 2.2×104 – 4.5×105 L mol−1sec−1 in seawater.[12] The dismutation of superoxide to hydrogen peroxide can also be catalyzed by the antioxidant enzyme superoxide dismutase with a rate constant on the order of 2×109 L mol−1sec−1.[32] As a result of these fast acting processes, the steady state concentration of superoxide is very small. Since superoxide is also moderately reactive towards trace metals and dissolved organic matter, any remaining superoxide is thought to be removed from the water column through reactions with these species.[12][30] As a result, the presence of superoxide in surface waters has been known to result in an increase of reduced iron.[30][33] This, in turn, serves to enhance the availability of iron to phytoplankton whose growth is often limited by this key nutrient. As a charged radical species, superoxide is unlikely to significantly affect an organism's cellular function since it is not able to easily diffuse through the cell membrane. Instead, its potential toxicity lies in its ability to react with extracellular surface proteins or carbohydrates to inactivate their functions.[34] Although its lifetime is fairly short (about 50 microseconds), superoxide has the potential to reach cell surfaces since it has a diffusion distance of about 320 nm.[1][34]

Hydrogen peroxide production

The reduction product of superoxide is hydrogen peroxide, one of the most studied reactive oxygen species because it occurs in relatively high concentrations, is relatively stable, and is fairly easy to measure.[12] It is thought that algal photosynthesis is one of the major modes of hydrogen peroxide production, while the production of H2O2 by stressed organisms is a secondary source.[13][14][15] In marine systems, hydrogen peroxide (H2O2) exists at concentrations of 10−8-10−9 M in the photic zone,[15] but has been found in double those concentrations in parts of the Atlantic Ocean.[35] Its lifetime, ranging from hours to days in coastal waters, can be as long as 15 days in Antarctic seawater.[12][30] H2O2 is important in aquatic environments because it can oxidize dissolved organic matter and affect the redox chemistry of iron, copper, and manganese.[33] Since hydrogen peroxide, as an uncharged molecule, diffuses easily across biological membranes it can directly damage cellular constituents (DNA and enzymes) by reacting with them and deactivating their functions.[2] In addition, hydrogen peroxide reduces to the hydroxyl radical, the most reactive radical and the one with the greatest possibility for damage.[1][2][12][30]

Hydroxyl radical production

Even though the superoxide and the hydrogen peroxide radicals are toxic in their own right, they become potentially more toxic when they interact to form the hydroxyl radical (OH•). This proceeds through the iron and copper catalyzed Haber–Weiss reaction:[36] O2 + Fe3+ ↔ O2 + Fe2+ H2O2 + Fe2+ ↔ Fe3+ + OH• + OH

Since iron and copper are present in coastal waters, the hydroxyl radical could be formed by reactions with either of the,[37][38] and, in fact, their oxidation does result in significant sources of hydroxyl radicals in the ocean.[33] The hydroxyl radical is the most unstable of the ROS (lifetime of 10−7seconds), reacting with many inorganic and organic species in the surrounding environment at rates near the diffusion limit (rate constants of 108 -1010 L mol−1 sec−1).[39] In seawater, the radical is removed as a result of reactions with bromide ions, while in fresh water it reacts principally with bicarbonate and carbonate ions.[12][30] Because it has such a high reactivity, day time concentrations in surface waters of the hydroxyl radical are generally very low (10−19 to 10−17 M).[12] The hydroxyl radical can oxidize membrane lipids and cause nucleic acids and proteins to denature. However, because the radical is so reactive, there is likely not enough time for transport to the cell surface (mean diffusion distance of 4.5 nm).[39] Thus, direct effects to organisms of externally generated hydroxyl radicals are expected to be minimal. Indirectly, the hydroxyl radical can result in significant biogeochemical changes in marine systems by influencing the cycling of dissolved organic matter and trace metal speciation. Both intracellular and extracellular reactive oxygen species can be removed from the environment by antioxidants produced biologically as a defense mechanism. Many phytoplankton, for instance, have been found to have numerous superoxide-scavenging (superoxide dismutase) and hydrogen peroxide-scavenging enzymes (catalase, ascorbate peroxidase, and glutathione peroxidase).[40][41][42][43][44] The antioxidant superoxide dismutase catalyses the formation of hydrogen peroxide from the superoxide anion through the following reaction:[45] 2 O2 + 2H+ ↔ O2 + H2O2. Similarly, catalase increases the formation of water from hydrogen peroxide by catalyzing the reaction:[46] 2H2O2↔ O2 + 2H2O. As a result of this reaction, the hydroxyl radical is prevented from forming. In addition, the presence of large quantities of humics in the water can also act as antioxidants of ROS.[47] However, it must be noted that certain ROS can inactivate certain enzymes. For instance, the superoxide anion is known to temporarily inhibit the function of catalase at high concentrations.[48]

Controls of ROS production in algae

Many algal species have been shown to not only produce reactive oxygen species under normal conditions but to increase production of these compounds under stressful situations. In particular, ROS levels have been shown to be influenced by cell size, cell density, growth stage, light intensity, temperature, and nutrient availability.

Cell size

Oda et al.[16] found that differences in the production of ROS were due to the size of the cell. By comparing four species of flagellates, they showed that the larger species Ichatonella produced the most superoxide and hydrogen peroxide per cell than Heterosigma akashiwo, Olisthodiscus luteus, and Fibrocapsa japonica. In a comparison of 37 species of marine microalgae, including dinoflagellates, rhaphidophytes, and chlorophytes, Marshall et al.[17] also found a direct relationship between cell size and the amount of superoxide produced. The largest cells, Chattonella marina, produced up to 100 times more superoxide than most other marine algae (see figure in [49]). The authors suggest that since ROS is produced as a byproduct of metabolism, and larger cells are more metabolically active than smaller cells, it follows that larger cells should produce more ROS. Similarly, since photosynthesis also produces ROS, larger cells likely have a greater volume of chloroplasts and would be expected to produce more ROS than smaller cells.

Algal density

The production of ROS has also been shown to be dependent on algal cell density. Marshall et al.[17] found that for Chattonella marina, higher concentrations of cells produced less superoxide per cell than those with a lower density. This may explain why some raphydophyte blooms are toxic at low concentration and non-toxic in heavy blooms.[50] Tang & Gobler[51] also found that cell density was inversely related to ROS production for the alga Cochlodinium polykrikoides. They found, in addition, that increases of ROS production were also related to the growth phase of algae. In particular, algae in exponential growth were more toxic than those in the stationary or late exponential phase. Many other algal species (Heterosigma akashiwo, Chattonella marina, and Chattonella antiqua) have also been shown to produce the highest amounts of ROS during the exponential phase of growth.[50][52] Oda et al.[16] suggest this is due to actively growing cells having higher photosynthesis and metabolic rates. Resting stage cells of Chattonella antiqua have been shown to generate less superoxide than their motile counterparts.[53]

Light levels

Since superoxide is produced through the auto-oxidation of an electron acceptor in photosystem I during photosynthesis, one would expect a positive relationship between light levels and algal ROS production.[17] This is indeed what has been shown: in the diatom Thallasia weissflogii, an increase in light intensity caused an increase in the production of both superoxide and hydrogen peroxide.[54] Similarly, in the flagellates Chattonella marina, Prorocentrum minimum, and Cochlodinium polykrikoides, decreases in light levels resulted in decreases in superoxide production,[17][55][56] with higher levels produced during the day. However, because many studies have found ROS production to be relatively high even in the dark, metabolic pathways other than photosynthesis are likely more important for production.[52] For instance, Liu et al.[57] found that ROS production was regulated by iron concentration and pH. From this evidence they suggest that ROS production is most likely due to a plasma membrane enzyme system dependent on iron availability. Similarly, in Heterosigma akashiwo, the depletion of iron and an increase in temperature, not light intensity, resulted in enhanced production of ROS.[50] Liu et al.[57] found the same relationship with temperature.

Functions of algal produced ROS

The active release of reactive oxygen species from cells has a variety of purposes, including a means to deter predators, or a chemical defense for the incapacitation of competitors.[58][59][60] In addition, ROS may be involved in cell signaling, as well as the oxidation or reduction of necessary or toxic metals.[13][61]

Chemical defense

It is not surprising that ROS production may be a form of chemical defense against predators, since at low levels it can damage DNA and at high levels lead to cell necrosis.[25] One of the most common mechanisms of cellular injury is the reaction of ROS with lipids, which can disrupt enzyme activity and ATP production, and lead to apoptosis.[37] Reactions of ROS with proteins can modify amino acids, fragment peptide chains, alter electrical charges, and ultimately inactivate an enzyme's function.[62][63] In DNA, deletions, mutations, and other lethal genetic effects may result from reactions with ROS.[64][65] Reactive oxygen species are especially inexpensive to produce as defense chemicals, simply because they are not composed of metabolically costly elements such as carbon, nitrogen, or phosphate. Reactive oxygen species produced by phytoplankton have been linked to deaths of fish, shellfish, and protists, as well as shown to reduce the viability and growth of bacteria.[20][50][66][67] In addition, a study by Marshall et al.[17] showed that four algal species used as bivalve feed produced significantly lower concentrations of superoxide, suggesting that ROS production by other algal species may be a way to decrease grazing by bivalves. The most direct evidence for ROS as a defense mechanism is the fact that many icthyotoxic algae produce greater concentrations of ROS than nonichthyotoxic strains.[16][17][19][50]

Enhancement of toxic exudates

It is possible that ROS may not be the actual toxic substance, but may in fact work to make other exudates more toxic by oxidizing them.[17][68] For instance, ROS from Chattonella marina have been shown to enhance the toxic effects of fatty acid eicosapentaenoic acid (EPA) on exposed fishes.[17][68] Similarly, free-fatty acids released from diatom biofilms as products of ROS oxidation of EPA are known to be toxic to zooplankters.[69] In addition, Fontana et al.[70] suggested that the interaction of ROS and diatom exudates (such as fatty acid hydroperoxides) are responsible for inhibiting embryonic development and causing larval abnormalities in copepods. Finally, ROS oxidation of algal polyunsaturated fatty acids have also been shown to deter grazers.[71]

Competitive advantage

In addition to impacting predator-prey interactions, the production of ROS may also help an alga get an advantage in the competition for resources against other algae, be a way to prevent fouling bacteria, and act as a signaling mechanism between cells.[60][67][72] ROS can inhibit photosynthesis in algae[25] Thus an alga that is more tolerant of ROS than another may produce and release it as a means of decreasing the other species competitive ability. In addition, Chattonella marina, the most well studied raphydophyte for ROS production, may produce a boundary of ROS that deters other marine microalgae from using nutrients in its vicinity.[27] Similarly, this boundary could also be a way to discourage bacteria fouling, since the production of ROS is known to inhibit growth and bioluminescent ability in the bacteria Vibrio alginolyticus and Vibrio fischeri, respectively.[67][72] Lastly, Marshall et al.[27] showed that Chattonella marina cells were able to change their rate of superoxide production in as little as one hour when in different cell densities, increasing the rate from 1.4 to 7.8 times the original. They suggest that this quick response in altering rates of production may be a form of chemical signaling between cells that works to provide information about cell density.

Reduction of metals

ROS may be useful in the oxidation or reduction of necessary or toxic metals. Since iron is necessary for phytoplankton growth, the auto-reduction of reactive oxygen species may be a way for algae to get usable iron from free or organically bound ferric iron.[73] For instance, Cakman et al.[74] showed that ROS may increase the amount of iron available through extracellular ferric reduction. It is thought that the high reducing power of this reaction is maintained through the electron-rich superoxide ion.[74] In several studies on the ROS production of Heterosigma akashiwo, hydrogen peroxide production was found to be inversely proportional to the concentration of iron available.[50][75] In addition, Cornish and Page in 1998 found that phytoplankton produce more ROS when there are lower levels of extracellular iron. They suggested that when intracellular iron is limiting, the phytoplankton respond by producing more ROS as a way to increase the reducing potential around the cell and thus be better able to reduce that iron to a usable form. Similarly, lower ROS production would suggest that the intracellular iron is at sufficiently high levels for cellular function.


References

  1. Cadenas, E (1989). "Biochemistry of oxygen toxicity". Annual Review of Biochemistry. 58: 79–110. doi:10.1146/annurev.bi.58.070189.000455. PMID 2673022.
  2. Fridovich, I (1998). "Oxygen toxicity: A radical explanation". The Journal of Experimental Biology. 201 (Pt 8): 1203–9. doi:10.1242/jeb.201.8.1203. PMID 9510531.
  3. Zheng, M.; Aslund, F; Storz, G (1998). "Activation of the OxyR Transcription Factor by Reversible Disulfide Bond Formation". Science. 279 (5357): 1718–21. Bibcode:1998Sci...279.1718Z. doi:10.1126/science.279.5357.1718. PMID 9497290.
  4. Martindale, JL; Holbrook, NJ (2002). "Cellular response to oxidative stress: Signaling for suicide and survival". Journal of Cellular Physiology. 192 (1): 1–15. doi:10.1002/jcp.10119. PMID 12115731.
  5. Harman, Denham (1992). "Free radical theory of aging". Mutation Research/DNAging. 275 (3–6): 257–266. doi:10.1016/0921-8734(92)90030-S. PMID 1383768.
  6. De Gara, Laura; De Pinto, Maria C.; Tommasi, Franca (2003). "The antioxidant systems vis-à-vis reactive oxygen species during plant–pathogen interaction". Plant Physiology and Biochemistry. 41 (10): 863–870. doi:10.1016/s0981-9428(03)00135-9.
  7. Doke, N (1985). "NADPH-dependent O2 generation in membrane fractions isolated from wounded potato tubers inoculated with Phytophthora infestans". Physiological Plant Pathology. 27 (3): 311–322. doi:10.1016/0048-4059(85)90044-X.
  8. Morel, Francoise; Doussiere, Jacques; Vignais, Pierre V. (1991). "The superoxide-generating oxidase of phagocytic cells. Physiological, molecular and pathological aspects". European Journal of Biochemistry. 201 (3): 523–46. doi:10.1111/j.1432-1033.1991.tb16312.x. PMID 1657601.
  9. Bolwell, GP; Butt, VS; Davies, DR; Zimmerlin, A (1995). "The origin of the oxidative burst in plants". Free Radical Research. 23 (6): 517–32. doi:10.3109/10715769509065273. PMID 8574346.
  10. Bolwell, G.P.; Wojtaszek, P. (1997). "Mechanisms for the generation of reactive oxygen species in plant defence – a broad perspective". Physiological and Molecular Plant Pathology. 51 (6): 347–366. doi:10.1006/pmpp.1997.0129.
  11. Dangl, Jeffery L.; Jones, Jonathan D. G. (2001). "Plant pathogens and integrated defence responses to infection". Nature. 411 (6839): 826–33. Bibcode:2001Natur.411..826D. doi:10.1038/35081161. PMID 11459065. S2CID 4345575.
  12. Kieber, David J.; Peake; Scully, Norman M. (2003). "Chapter 8: Reactive oxygen species in aquatic ecosystems". In E. Walter Helbling (ed.). UV effects in aquatic organisms and ecosystems. Cambridge: Royal Society of Chemistry. pp. 251–76. ISBN 9780854043019.
  13. Palenik, Brian; Zafiriou, O. C.; Morel, F. M. M. (1987). "Hydrogen Peroxide Production by a Marine Phytoplankter". Limnology and Oceanography. 32 (6). American Society of Limnology and Oceanography: 1365–1369. Bibcode:1987LimOc..32.1365P. doi:10.4319/lo.1987.32.6.1365. JSTOR 2836931.
  14. Palenik, B.; Morel, F. M. M. (1988). "Dark production of H2O2 in the Sargasso Sea". Limnology and Oceanography. 33 (6, part 2): 1606–11. doi:10.4319/lo.1988.33.6_part_2.1606.
  15. Wong, George T.F.; Dunstan, William M.; Kim, Dong-Beom (2003). "The decomposition of hydrogen peroxide by marine phytoplankton". Oceanologica Acta. 26 (2): 191–198. doi:10.1016/S0399-1784(02)00006-3.
  16. Oda, T.; Nakamura, A.; Shikayama, M.; Kawano, I.; Ishimatsu, A.; Muramatsu, T. (1997). "Generation of reactive oxygen species by raphidophycean phytoplankton". Bioscience, Biotechnology, and Biochemistry. 61 (10): 1658–62. doi:10.1271/bbb.61.1658. PMID 9362113.
  17. Ishimatsu, Atsushi; Oda, Tatsuya; Yoshida, Makoto; Ozaki, Masayori (1996). "Oxygen radicals are probably involved in the mortality of Yellowtail by Chattonella marinas". Fisheries Science. 62 (5): 836–837. doi:10.2331/fishsci.62.836.
  18. Evans, Claire; Malin, Gillian; Mills, Graham P.; Wilson, William H. (2006). "Viral Infection of Emiliania Huxleyi (Prymnesiophyceae) Leads to Elevated Production of Reactive Oxygen Species". Journal of Phycology. 42 (5): 1040–47. doi:10.1111/j.1529-8817.2006.00256.x. S2CID 84678114.
  19. Flores, HS; Wikfors, GH; Dam, HG (2012). "Reactive oxygen species are linked to the toxicity of the dinoflagellate Alexandrium spp. To protists". Aquatic Microbial Ecology. 66 (2): 199–209. doi:10.3354/ame01570. hdl:1912/27247.
  20. Fan, Song-Miao (2008). "Photochemical and biochemical controls on reactive oxygen and iron speciation in the pelagic surface ocean". Marine Chemistry. 109 (1–2): 152–164. doi:10.1016/j.marchem.2008.01.005.
  21. Moller, IM (2001). "Plant Mitochondria and Oxidative Stress: Electron Transport, NADPH Turnover, and Metabolism of Reactive Oxygen Species". Annual Review of Plant Physiology and Plant Molecular Biology. 52: 561–591. doi:10.1146/annurev.arplant.52.1.561. PMID 11337409.
  22. Lesser, MP (2006). "Oxidative stress in marine environments: Biochemistry and physiological ecology". Annual Review of Physiology. 68: 253–78. doi:10.1146/annurev.physiol.68.040104.110001. PMID 16460273.
  23. Corpas, FJ; Barroso, JB; Del Río, LA (2001). "Peroxisomes as a source of reactive oxygen species and nitric oxide signal molecules in plant cells". Trends in Plant Science. 6 (4): 145–50. doi:10.1016/S1360-1385(01)01898-2. PMID 11286918.
  24. Marshall, Judith-Anne; Salas, Miguel; Oda, Tatsuya; Hallegraeff, Gustaaf (2005). "Superoxide production by marine microalgae". Marine Biology. 147 (2): 533–540. doi:10.1007/s00227-005-1596-7. S2CID 82978185.
  25. Jamieson, D; Chance, B; Cadenas, E; Boveris, A (1986). "The relation of free radical production to hyperoxia". Annual Review of Physiology. 48: 703–19. doi:10.1146/annurev.ph.48.030186.003415. PMID 3010832.
  26. Land, Edward J.; Swallow, Albert J. (1969). "One-electron reactions in biochemical systems as studied by pulse radiolysis. II. Riboflavine". Biochemistry. 8 (5): 2117–25. doi:10.1021/bi00833a050. PMID 5785230.
  27. Blough, Neil; Zepp, Richard G. (1995). "Chapter 8: Reactive oxygen species in natural waters". In Christopher S. Foote (ed.). Active oxygen in chemistry. London: Blackie Acad. & Professional. ISBN 9780751402926.
  28. Kustka, Adam B.; Shaked, Yeala; Milligan, Allen J.; King, D. Whitney; Morel, François M. M. (2005). "Extracellular production of superoxide by marine diatoms: Contrasting effects on iron redox chemistry and bioavailability". Limnology and Oceanography. 50 (4): 1172–80. Bibcode:2005LimOc..50.1172K. doi:10.4319/lo.2005.50.4.1172.
  29. Moffett, James W.; Zika, Rod G. (1987). "Reaction kinetics of hydrogen peroxide with copper and iron in seawater". Environmental Science & Technology. 21 (8): 804–810. Bibcode:1987EnST...21..804M. doi:10.1021/es00162a012. PMID 19995065.
  30. Fridovich, I (1986). "Biological effects of the superoxide radical". Archives of Biochemistry and Biophysics. 247 (1): 1–11. doi:10.1016/0003-9861(86)90526-6. PMID 3010872.
  31. Yuan, Jinchun; Shiller, Alan M (2001). "The distribution of hydrogen peroxide in the southern and central Atlantic ocean". Deep-Sea Research Part II: Topical Studies in Oceanography. 48 (13): 2947–2970. Bibcode:2001DSRII..48.2947Y. doi:10.1016/S0967-0645(01)00026-1.
  32. Klebanoff, S.J. (1980). Furth, Ralph van (ed.). Mononuclear phagocytes: functional aspects. Boston: Martinus Nijhoff. pp. 1105–1141. ISBN 9789024722112.
  33. Halliwell, Barry; Gutteridge, John M. C. (1999). Free Radical Biology and Medicine (3rd ed.). Oxford: Clarendon Press. ISBN 9780198500452.
  34. Oda, Tatsuya; Moritomi, Junko; Kawano, Ienobu; Hamaguchi, Shiho; Ishimatsu, Atsushi; Muramatsu, Tsuyosi (1995). "Catalase- and Superoxide Dismutase-induced Morphological Changes and Growth Inhibition in the Red Tide Phytoplankton Chattonella marina". Bioscience, Biotechnology, and Biochemistry. 59 (11): 2044–48. doi:10.1271/bbb.59.2044.
  35. Simpson, JA; Cheeseman, KH; Smith, SE; Dean, RT (1988). "Free-radical generation by copper ions and hydrogen peroxide. Stimulation by Hepes buffer". The Biochemical Journal. 254 (2): 519–23. doi:10.1042/bj2540519. PMC 1135108. PMID 3178771.
  36. K., Asada; Baker, N.R.; Boyer, J.B. (1994). "Mechanisms for scavenging reactive molecules generated in chloroplasts under light stress". Photoinhibition of photosynthesis: from molecular mechanisms to the field. Oxford, UK: Bios Scientific Publishers. pp. 129–142. ISBN 9781872748030.
  37. Collén, Jonas; Pedersén, Marianne (1996). "Production, scavenging and toxicity of hydrogen peroxide in the green seaweed Ulva rigida". European Journal of Phycology. 31 (3): 265–271. doi:10.1080/09670269600651471.
  38. Barros, MP; Granbom, M; Colepicolo, P; Pedersén, M (2003). "Temporal mismatch between induction of superoxide dismutase and ascorbate peroxidase correlates with high H2O2 concentration in seawater from clofibrate-treated red algae Kappaphycus alvarezii". Archives of Biochemistry and Biophysics. 420 (1): 161–8. doi:10.1016/j.abb.2003.09.014. PMID 14622986.
  39. Dummermuth, A.L; Karsten, U; Fisch, K.M; König, G.M; Wiencke, C (2003). "Responses of marine macroalgae to hydrogen-peroxide stress" (PDF). Journal of Experimental Marine Biology and Ecology. 289: 103–121. doi:10.1016/S0022-0981(03)00042-X.
  40. Choo, Kyung-sil; Snoeijs, Pauli; Pedersén, Marianne (2004). "Oxidative stress tolerance in the filamentous green algae Cladophora glomerata and Enteromorpha ahlneriana". Journal of Experimental Marine Biology and Ecology. 298: 111–123. doi:10.1016/j.jembe.2003.08.007.
  41. Bannister, JV; Bannister, WH; Rotilio, G (1987). "Aspects of the structure, function, and applications of superoxide dismutase". Critical Reviews in Biochemistry. 22 (2): 111–80. doi:10.3109/10409238709083738. PMID 3315461.
  42. Forman, H.; Fisher, A.B. (1981). "Antioxidant defense". In Gilbert, Daniel L. (ed.). Oxygen and living processes: An interdisciplinary approach. New York: Springer-Verlag. ISBN 978-0387905549.
  43. Sandvik, Sonya L.Holder; Bilski, Piotr; Pakulski, J.Dean; Chignell, Colin F.; Coffin, Richard B. (2000). "Photogeneration of singlet oxygen and free radicals in dissolved organic matter isolated from the Mississippi and Atchafalaya River plumes". Marine Chemistry. 69 (1–2): 139–152. doi:10.1016/S0304-4203(99)00101-2.
  44. Singleton, Paul; Sainsbury, Diana (2001). Dictionary or microbiology and molecular biology (3rd ed.). Chichester, UK: Wiley. p. 751. ISBN 978-0-470-03545-0.
  45. Mooney, BD; Dorantes-Aranda, JJ; Place, AR; Hallegraeff, GM (2011). "Ichthyotoxicity of gymnodinioid dinoflagellates: PUFA and superoxide effects in sheepshead minnow larvae and rainbow trout gill cells" (PDF). Marine Ecology Progress Series. 426: 213–224. Bibcode:2011MEPS..426..213M. doi:10.3354/meps09036.
  46. Tang, Ying Zhong; Gobler, Christopher J. (2009). "Characterization of the toxicity of Cochlodinium polykrikoides isolates from Northeast US estuaries to finfish and shellfish". Harmful Algae. 8 (3): 454–62. doi:10.1016/j.hal.2008.10.001.
  47. Portune, Kevin J.; Craig Cary, Stephen; Warner, Mark E. (2010). "Antioxidant Enzyme Response and Reactive Oxygen Species Production in Marine Raphidophytes1". Journal of Phycology. 46 (6): 1161–1171. doi:10.1111/j.1529-8817.2010.00906.x. S2CID 83834408.
  48. Tanaka, Kazuko; Muto, Yoshinori; Shimada, Masahisa (1994). "Generation of superoxide anion radicals by the marine phytoplankton organism, Chattonella antiqua". Journal of Plankton Research. 16 (2): 161–69. doi:10.1093/plankt/16.2.161.
  49. Milne, Angela; Davey, Margaret S.; Worsfold, Paul J.; Achterberg, Eric P.; Taylor, Alison R. (2009). "Real-time detection of reactive oxygen species generation by marine phytoplankton using flow injection—chemiluminescence" (PDF). Limnology and Oceanography: Methods. 7 (10): 706–15. doi:10.4319/lom.2009.7.706. S2CID 59438207.
  50. d(-k.)., Kim; t., Okamoto; t., Oda; k., Tachibana; k.s., Lee; a., Ishimatsu; y., Matsuyama; t., Honjo; t., Muramatsu (2001). "Possible involvement of the glycocalyx in the ichthyotoxicity of Chattonella marina (Raphidophyceae): Immunological approach using antiserum against cell surface structures of the flagellate". Marine Biology. 139 (4): 625–632. doi:10.1007/s002270100614. S2CID 84678064.
  51. Park, So Yun; Choi, Eun Seok; Hwang, Jinik; Kim, Donggiun; Ryu, Tae Kwon; Lee, Taek-Kyun (2010). "Physiological and biochemical responses of Prorocentrum minimum to high light stress". Ocean Science Journal. 44 (4): 199–204. Bibcode:2009OSJ....44..199P. doi:10.1007/s12601-009-0018-z. S2CID 83931880.
  52. Liu, Wenhua; Au, Doris W.T.; Anderson, Donald M.; Lam, Paul K.S.; Wu, Rudolf S.S. (2007). "Effects of nutrients, salinity, pH and light:dark cycle on the production of reactive oxygen species in the alga Chattonella marina" (PDF). Journal of Experimental Marine Biology and Ecology. 346 (1–2): 76–86. doi:10.1016/j.jembe.2007.03.007. hdl:1912/1764.
  53. Cembella, A.D; Quilliam, M.A; Lewis, N.I; Bauder, A.G; Dell'Aversano, C; Thomas, K; Jellett, J; Cusack, R.R (2002). "The toxigenic marine dinoflagellate Alexandrium tamarense as the probable cause of mortality of caged salmon in Nova Scotia". Harmful Algae. 1 (3): 313–325. doi:10.1016/S1568-9883(02)00048-3.
  54. Legrand, Catherine; Rengefors, Karin; Fistarol, Giovana O.; Granéli, Edna (2003). "Allelopathy in phytoplankton - biochemical, ecological and evolutionary aspects". Phycologia. 42 (4): 406–419. doi:10.2216/i0031-8884-42-4-406.1. S2CID 84166335.
  55. Granéli, E.; Hansen, P. J. (2006). "Allelopathy in Harmful Algae: A Mechanism to Compete for Resources?". Ecology of Harmful Algae. Ecological Studies. Vol. 189. Springer. pp. 189–206. doi:10.1007/978-3-540-32210-8_15. ISBN 978-3-540-32209-2.
  56. Stadtman, E.R. (1986). "Oxidation of proteins by mixed-function oxidation systems: Implication in protein turnover, ageing and neutrophil function". Trends in Biochemical Sciences. 11: 11–12. doi:10.1016/0968-0004(86)90221-5.
  57. Davies, KJ (1987). "Protein damage and degradation by oxygen radicals. I. General aspects". The Journal of Biological Chemistry. 262 (20): 9895–901. doi:10.1016/S0021-9258(18)48018-0. PMID 3036875.
  58. Imlay, JA; Linn, S (1988). "DNA damage and oxygen radical toxicity". Science. 240 (4857): 1302–9. Bibcode:1988Sci...240.1302I. doi:10.1126/science.3287616. PMID 3287616.
  59. Imlay, JA (2003). "Pathways of oxidative damage". Annual Review of Microbiology. 57: 395–418. doi:10.1146/annurev.micro.57.030502.090938. PMID 14527285.
  60. Babior, BM (1978). "Oxygen-dependent microbial killing by phagocytes (first of two parts)". The New England Journal of Medicine. 298 (12): 659–68. doi:10.1056/NEJM197803232981205. PMID 24176.
  61. Kim, Daekyung; Nakamura, Atsushi; Okamoto, Tarou; Komatsu, Nobukazu; Oda, Tatsuya; Ishimatsu, Atsushi; Muramatsu, Tsuyoshi (1999). "Toxic potential of the raphidophyte Olisthodiscus luteus: mediation by reactive oxygen species". Journal of Plankton Research. 21 (6): 1017–27. doi:10.1093/plankt/21.6.1017. ISSN 0142-7873.
  62. Okaichi, Tomotoshi; Mark Anderson, Donald; Nemoto, Takahisa (1989). "Proceedings of the first international symposium on Red tides". Red tides: biology, environmental science, and toxicology. New York: Elsevier. pp. 145–176. ISBN 9780444013439.
  63. Juttner, Friedrich (2001). "Liberation of 5,8,11,14,17-Eicosapentaenoic Acid and Other Polyunsaturated Fatty Acids from Lipids As a Grazer Defense Reaction in Epilithic Diatom Biofilms". Journal of Phycology. 37 (5): 744–55. doi:10.1046/j.1529-8817.2001.00130.x. S2CID 86143120.
  64. Fontana, A; d'Ippolito, G; Cutignano, A; Romano, G; Lamari, N; Massa Gallucci, A; Cimino, G; Miralto, A; Ianora, A (2007). "LOX-induced lipid peroxidation mechanism responsible for the detrimental effect of marine diatoms on zooplankton grazers". ChemBioChem. 8 (15): 1810–8. doi:10.1002/cbic.200700269. PMID 17886321. S2CID 808518.
  65. Ikawa, M. (2004). "Algal polyunsaturated fatty acids and effects on plankton ecology and other organisms" (PDF). UNH Center for Freshwater Biology Research. 6 (2): 17–44.
  66. Oda, T.; Ishimatsu, A.; Shimada, M.; Takeshita, S.; Muramatsu, T. (1992). "Oxygen-radical-mediated toxic effects of the red tide flagellate Chattonella marina on Vibrio alginolyticus". Marine Biology. 112 (3): 505–9. doi:10.1007/BF00356297. S2CID 82953809.
  67. Sunda, William G.; Huntsman, Susan A. (1995). "Iron uptake and growth limitation in oceanic and coastal phytoplankton". Marine Chemistry. 50 (1–4): 189–206. doi:10.1016/0304-4203(95)00035-p.
  68. Cakmak, I; Van De Wetering, DA; Marschner, H; Bienfait, HF (1987). "Involvement of superoxide radical in extracellular ferric reduction by iron-deficient bean roots". Plant Physiology. 85 (1): 310–4. doi:10.1104/pp.85.1.310. PMC 1054247. PMID 16665677.
  69. Twiner, Michael J.; Dixon, S. Jeffrey; Trick, Charles G. (2001). "Toxic effects of Heterosigma akashiwo do not appear to be mediated by hydrogen peroxide". Limnology and Oceanography. 46 (6): 1400–05. Bibcode:2001LimOc..46.1400T. doi:10.4319/lo.2001.46.6.1400.

Share this article:

This article uses material from the Wikipedia article Reactive_oxygen_species_production_in_marine_microalgae, and is written by contributors. Text is available under a CC BY-SA 4.0 International License; additional terms may apply. Images, videos and audio are available under their respective licenses.