Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Saxitoxin and tetrodotoxin bioavailability increases in future oceans

Abstract

Increasing atmospheric CO2 levels are largely absorbed by the ocean, decreasing surface water pH1. In combination with increasing ocean temperatures, these changes have been identified as a major sustainability threat to future marine life2. Interactions between marine organisms are known to depend on biomolecules, although the influence of oceanic pH on their bioavailability and functionality remains unexplored. Here we show that global change substantially impacts two ecological keystone molecules3 in the ocean, the paralytic neurotoxins saxitoxin and tetrodotoxin. Increasing temperatures and declining pH increase the abundance of their toxic forms in the water. Our geospatial global model predicts where this increased toxicity could intensify the devastating impact of harmful algal blooms, for example through an increased incidence of paralytic shellfish poisoning. Calculations of future saxitoxin toxicity levels in Alaskan clams, Saxidomus gigantea, show critical exceedance of limits safe for consumption. Our findings for saxitoxin and tetrodotoxin exemplify potential consequences of changing pH and temperature on chemicals dissolved in the sea. This reveals major implications not only for ecotoxicology, but also for chemical signals that mediate species interactions such as foraging, reproduction or predation in the ocean, with unexplored consequences for ecosystem stability and ecosystem services.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Neurotoxin structures, charge distributions and relative proportions of individual protonation states.
Fig. 2: Global abundance of the toxic saxitoxin protonation state today and in the future.
Fig. 3: Amount of the toxic saxitoxin form present in butter clam Saxidomus gigantea tissue.

Similar content being viewed by others

Data availability

Source data for curves in Fig. 1 calculated based on the given references and coordinates of the molecular structures are available from the corresponding author upon request. The data used to generate the dataset for Fig. 2 are available from the Harmful Algal Information System metadatabase (HAEDAT, http://haedat.iode.org), the NOAA COPEPOD database (https://www.st.nmfs.noaa.gov/copepod), the Global marine environment dataset (GMED, http://gmed.auckland.ac.nz) and the IPCC (WCRP CMPI3) multi-model database (https://cmip.llnl.gov). Data used to generate Fig. 3 can be accessed via the website of the Qagan Tayagungin Tribe (https://www.qttribe.org>Environment>PSP Program). The extracted, collated data supporting the findings in our study are deposited in the PANGAEA archive and available at https://doi.org/10.1594/PANGAEA.904260.

Code availability

The code used to calculate the proportions of different protonation states is available from the corresponding author on request .

References

  1. IPCC Climate Change 2014: Synthesis Report (eds Core Writing Team, Pachauri, R. K. & Meyer L. A.) (IPCC, 2014).

  2. DeWeerdt, S. Sea change. Nature 550, S54–S58 (2017).

    Article  CAS  Google Scholar 

  3. Ferrer, R. P. & Zimmer, R. K. Molecules of keystone significance: crucial agents in ecology and resource management. Bioscience 63, 428–438 (2013).

    Article  Google Scholar 

  4. Caldeira, K. & Wickett, M. E. Oceanography: anthropogenic carbon and ocean pH. Nature 425, 365 (2003).

    Article  CAS  Google Scholar 

  5. Baumann, H. & Smith, E. M. Quantifying metabolically driven pH and oxygen fluctuations in US nearshore habitats at diel to interannual time scales. Estuaries Coasts 41, 1102–1117 (2018).

    Article  CAS  Google Scholar 

  6. Fabry, V. J., Seibel, B. A., Feely, R. A. & Orr, J. C. Impacts of ocean acidification on marine fauna and ecosystem processes. ICES J. Mar. Sci. 65, 414–432 (2008).

    Article  CAS  Google Scholar 

  7. Wittmann, A. C. & Pörtner, H.-O. Sensitivities of extant animal taxa to ocean acidification. Nat. Clim. Change 3, 995–1001 (2013).

    Article  CAS  Google Scholar 

  8. Clements, J. C. & Hunt, H. L. Marine animal behaviour in a high CO2 ocean. Mar. Ecol. Prog. Ser. 536, 259–279 (2015).

    Article  CAS  Google Scholar 

  9. Roggatz, C. C., Lorch, M., Hardege, J. D. & Benoit, D. M. Ocean acidification affects marine chemical communication by changing structure and function of peptide signalling molecules. Glob. Change Biol. 22, 3914–3926 (2016).

    Article  Google Scholar 

  10. Porteus, C. S. et al. Near-future CO2 levels impair the olfactory system of a marine fish. Nat. Clim. Change 8, 737–743 (2018).

    Article  CAS  Google Scholar 

  11. Hay, M. E. Marine chemical ecology: chemical signals and cues structure marine populations, communities, and ecosystems. Ann. Rev. Mar. Sci. 1, 193–212 (2009).

    Article  Google Scholar 

  12. Bane, V., Lehane, M., Dikshit, M., O’Riordan, A. & Furey, A. Tetrodotoxin: chemistry, toxicity, source, distribution and detection. Toxins 6, 693–755 (2014).

    Article  CAS  Google Scholar 

  13. Cusick, K. D. & Sayler, G. S. An overview on the marine neurotoxin, saxitoxin: genetics, molecular targets, methods of detection and ecological functions. Mar. Drugs 11, 991–1018 (2013).

    Article  CAS  Google Scholar 

  14. Williams, B. L. Behavioral and chemical ecology of marine organisms with respect to tetrodotoxin. Mar. Drugs 8, 381–398 (2010).

    Article  CAS  Google Scholar 

  15. Wyatt, T. & Jenkinson, I. R. Notes on Alexandrium population dynamics. J. Plankton Res. 19, 551–575 (1997).

    Article  Google Scholar 

  16. Lefebvre, K. A. et al. Characterization of intracellular and extracellular saxitoxin levels in both field and cultured Alexandrium spp. samples from Sequim Bay, Washington. Mar. Drugs 6, 103–116 (2008).

    Article  CAS  Google Scholar 

  17. Gobler, C. J. et al. Ocean warming since 1982 has expanded the niche of toxic algal blooms in the North Atlantic and North Pacific oceans. Proc. Natl Acad. Sci. USA 114, 4975–4980 (2017).

    Article  CAS  Google Scholar 

  18. Riebesell, U. et al. Toxic algal bloom induced by ocean acidification disrupts the pelagic food web. Nat. Clim. Change 8, 1082–1086 (2018).

    Article  CAS  Google Scholar 

  19. Li, A., Chen, H., Qiu, J., Lin, H. & Gu, H. Determination of multiple toxins in whelk and clam samples collected from the Chukchi and Bering seas. Toxicon 109, 84–93 (2016).

    Article  CAS  Google Scholar 

  20. Starr, M. et al. Multispecies mass mortality of marine fauna linked to a toxic dinoflagellate bloom. PLoS ONE 12, e0176299 (2017).

    Article  CAS  Google Scholar 

  21. Wang, D.-Z. Neurotoxins from marine dinoflagellates: a brief review. Mar. Drugs 6, 349–371 (2008).

    Article  Google Scholar 

  22. Woodward, R. B. The structure of tetrodotoxin. J. Macromol. Sci. Pure Appl. Chem. 9, 49–74 (1964).

    Article  CAS  Google Scholar 

  23. Rogers, R. S. & Rapoport, H. The pK as of saxitoxin. J. Am. Chem. Soc. 102, 7335–7339 (1980).

    Article  CAS  Google Scholar 

  24. Hegyi, B. et al. Tetrodotoxin blockade on canine cardiac L-type Ca2+ channels depends on pH and redox potential. Mar. Drugs 11, 2140–2153 (2013).

    Article  CAS  Google Scholar 

  25. Yotsu-Yamashita, M., Sugimoto, A., Takai, A. & Yasumoto, T. Effects of specific modifications of several hydroxyls of tetrodotoxin on its affinity to rat brain membrane. J. Pharmacol. Exp. Ther. 289, 1688–1696 (1999).

    CAS  Google Scholar 

  26. Roggatz, C. C., Lorch, M. & Benoit, D. M. Influence of solvent representation on nuclear shielding calculations of protonation states of small biological molecules. J. Chem. Theory Comput. 14, 2684–2695 (2018).

    Article  CAS  Google Scholar 

  27. Colquhoun, D., Henderson, R. & Ritchie, J. M. The binding of labelled tetrodotoxin to non-myelinated nerve fibres. J. Physiol. 227, 95–126 (1972).

    Article  CAS  Google Scholar 

  28. Ulbricht, W. & Wagner, H. H. The influence of pH on the rate of tetrodotoxin action on myelinated nerve fibres. J. Physiol. 252, 185–202 (1975).

    Article  CAS  Google Scholar 

  29. Natsuike, M. et al. Possible spreading of toxic Alexandrium tamarense blooms on the Chukchi Sea shelf with the inflow of Pacific summer water due to climatic warming. Harmful Algae 61, 80–86 (2017).

    Article  Google Scholar 

  30. Blackburn, S. I., Hallegraeff, G. M. & Bolch, C. J. Vegetative reproduction and sexual life cycle of the toxic dinoflagellate Gymnodinium catenatum from Tasmania, Australia. J. Phycol. 25, 577–590 (1989).

    Article  Google Scholar 

  31. Hallegraeff, G. M. Ocean climate change, phytoplankton community responses, and harmful algal blooms: a formidable predictive challenge. J. Phycol. 46, 220–235 (2010).

    Article  CAS  Google Scholar 

  32. Hallegraeff, G. M. A review of harmful algal blooms and their apparent global increase. Phycologia 32, 79–99 (1993).

    Article  Google Scholar 

  33. Tatters, A. O., Flewelling, L. J., Fu, F., Granholm, A. A. & Hutchins, D. A. High CO2 promotes the production of paralytic shellfish poisoning toxins by Alexandrium catenella from Southern California waters. Harmful Algae 30, 37–43 (2013).

    Article  CAS  Google Scholar 

  34. Cembella, A. in Physiological Ecology of Harmful Algal Blooms (eds. Anderson, D. M. et al.) 381–404 (NATO ASI Series G 41, Springer, 1998).

  35. Lefebvre, K. A. et al. Prevalence of algal toxins in Alaskan marine mammals foraging in a changing Arctic and Subarctic environment. Harmful Algae 55, 13–24 (2016).

    Article  CAS  Google Scholar 

  36. Zonneveld, K. A. F. et al. Atlas of modern dinoflagellate cyst distribution based on 2405 data points. Rev. Palaeobot. Palynol. 191, 1–197 (2013).

    Article  Google Scholar 

  37. Bezençon, J. et al. pK a determination by 1H NMR spectroscopy—an old methodology revisited. J. Pharm. Biomed. Anal. 93, 147–155 (2014).

    Article  CAS  Google Scholar 

  38. Goto, T., Kishi, Y., Takahashi, S. & Hirata, Y. Tetrodotoxin. Tetrahedron 21, 2059–2088 (1965).

    Article  CAS  Google Scholar 

  39. Shimizu, Y. et al. Structure of neosaxitoxin. J. Am. Chem. Soc. 100, 6791–6793 (1978).

    Article  CAS  Google Scholar 

  40. Po, H. N. & Senozan, N. M. The Henderson–Hasselbalch equation: its history and limitations. J. Chem. Educ. 78, 1499 (2001).

    Article  CAS  Google Scholar 

  41. Reijenga, J. C., Gagliardi, L. G. & Kenndler, E. Temperature dependence of acidity constants, a tool to affect separation selectivity in capillary electrophoresis. J. Chromatogr. A 1155, 142–145 (2007).

    Article  CAS  Google Scholar 

  42. Global Climate Report—Annual 2018 State of the Climate (National Centers for Environmental Information, 2019).

  43. Kim, S. et al. PubChem substance and compound databases. Nucleic Acids Res. 44, D1202–D1213 (2016).

    Article  CAS  Google Scholar 

  44. Mosher, H. S. The chemistry of tetrodotoxin. Ann. N. Y. Acad. Sci. 479, 32–43 (1986).

    Article  CAS  Google Scholar 

  45. Adamo, C. & Barone, V. Toward reliable density functional methods without adjustable parameters: the PBE0 model. J. Chem. Phys. 110, 6158–6170 (1999).

    Article  CAS  Google Scholar 

  46. Jensen, F. Polarization consistent basis sets: principles. J. Chem. Phys. 115, 9113–9125 (2001).

    Article  CAS  Google Scholar 

  47. Jensen, F. Polarization consistent basis sets. II. Estimating the Kohn–Sham basis set limit. J. Chem. Phys. 116, 7372–7379 (2002).

    Article  CAS  Google Scholar 

  48. Jensen, F. Polarization consistent basis sets. III. The importance of diffuse functions. J. Chem. Phys. 117, 9234–9240 (2002).

    Article  CAS  Google Scholar 

  49. Klamt, A. Conductor-like screening model for real solvents: a new approach to the quantitative calculation of solvation phenomena. J. Phys. Chem. 99, 2224–2235 (1995).

    Article  CAS  Google Scholar 

  50. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2, 73–78 (2012).

    Article  CAS  Google Scholar 

  51. Neese, F., Wennmohs, F., Hansen, A. & Becker, U. Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys. 356, 98–109 (2009).

    Article  CAS  Google Scholar 

  52. Weigend, F. & Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 7, 3297–3305 (2005).

    Article  CAS  Google Scholar 

  53. Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H–Pu. J. Chem. Phys. 132, 154104 (2010).

    Article  CAS  Google Scholar 

  54. Bode, B. M. & Gordon, M. S. MacMolPlt: a graphical user interface for GAMESS. J. Mol. Graph. Model. 16, 164 (1998).

    Article  Google Scholar 

  55. Basher, Z., Costello, M. J. & Bowden, D. A. Global Marine Environment Dataset (GMED) (Univ. Auckland, 2015).

  56. Boyer, T. P. et al. World Ocean Database 2013 (National Centers for Environmental Information, 2013); https://www.nodc.noaa.gov/OC5/WOD13/

  57. Tyberghein, L. et al. Bio-ORACLE: a global environmental dataset for marine species distribution modelling. Glob. Ecol. Biogeogr. 21, 272–281 (2012).

    Article  Google Scholar 

  58. Gattuso, J.-P. et al. Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios. Science 349, aac4722 (2015).

    Article  CAS  Google Scholar 

  59. World Ocean Basemap (ArcGIS REST Services Directory, ESRI, accessed 28 May 2018); http://services.arcgisonline.com/arcgis/rest/services/Ocean/World_Ocean_Base/MapServer

  60. PSP Program (Quagan Tayagungin Tribe, accessed 14 August 2018); https://go.nature.com/2kXEn85

  61. Ringwood, A. H. & Keppler, C. J. Water quality variation and clam growth: Is pH really a non-issue in estuaries? Estuaries 25, 901–907 (2002).

    Article  Google Scholar 

  62. Booth, C. E., McDonald, D. G. & Walsh, P. J. Acid-base balance in the sea mussel, Mytilus edulis. I. Effects of hypoxia and air-exposure on hemolymph acid-base status. Mar. Biol. Lett. 5, 347–358 (1984).

    CAS  Google Scholar 

  63. Dwyer, J. J. III & Burnett, L. E. Acid-base status of the oyster Crassostrea virginica in response to air exposure and to infections by Perkinsus marinus. Biol. Bull. 190, 139–147 (1996).

    Article  Google Scholar 

  64. Fish and Fishery Products Hazards and Controls Guidance (Department of Health and Human Services, Public Health Service, Food and Drug Administration, Center for Food Safety and Applied Nutrition, Office for Food Safety, 2011).

Download references

Acknowledgements

We acknowledge the Viper High Performance Computing facility of the University of Hull and its support team. C.C.R. was funded through D. Parsons’ project grant (no. ERC-2016-COG GEOSTICK). We thank the Quagan Tayagungin Tribe for access to the clam toxicity data at the PSP Program website. We acknowledge the World Climate Research Programme’s Working Group on Coupled Modelling and the climate modelling groups for producing and making available their model output. For CMIP the US Department of Energy’s Program for Climate Model Diagnosis and Inter-comparison provides coordinating support and led the development of software infrastructure in partnership with the Global Organization for Earth System Science Portals. We thank D. Parsons, Energy and Environment Institute, University of Hull, and H. Bartels-Hardege, Biological and Marine Sciences, University of Hull, for valuable suggestions and discussions.

Author information

Authors and Affiliations

Authors

Contributions

C.C.R. and J.D.H. designed the study. K.C.W.V. and A.C.A. performed the analysis of the datasets for geospatial models and C.C.R. and D.M.B. calculated the molecular models. Access to toxin mussel data was provided by A.D. and B.W. C.C.R, J.D.H. and K.C.W.V. shared responsibility for, and N.F. contributed to, the writing of the manuscript. All authors contributed to the final version of the manuscript.

Corresponding author

Correspondence to C. C. Roggatz.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Climate Change thanks K. Cusick and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary Information

Supplementary methods, Figs. 1–3, Table 1 and references.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Roggatz, C.C., Fletcher, N., Benoit, D.M. et al. Saxitoxin and tetrodotoxin bioavailability increases in future oceans. Nat. Clim. Chang. 9, 840–844 (2019). https://doi.org/10.1038/s41558-019-0589-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41558-019-0589-3

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing