The spring bloom is a strong increase in phytoplankton abundance (i.e. stock) that typically occurs in the early spring and lasts until late spring or early summer. This seasonal event is characteristic of temperate North Atlantic, sub-polar, and coastal waters.[1][2] Phytoplankton blooms occur when growth exceeds losses, however there is no universally accepted definition of the magnitude of change or the threshold of abundance that constitutes a bloom. The magnitude, spatial extent and duration of a bloom depends on a variety of abiotic and biotic factors. Abiotic factors include light availability, nutrients, temperature, physical processes that influence light availability,[1][2][3][4][5] and biotic factors include grazing, viral lysis, and phytoplankton physiology.[6] The factors that lead to bloom initiation are still actively debated (see Critical Depth).

A spring bloom off the coast of Boracay Island in the Philippines






Classical mechanism edit

In the spring, more light becomes available and stratification of the water column occurs as increasing temperatures warm the surface waters (referred to as thermal stratification). As a result, vertical mixing is inhibited and phytoplankton and nutrients are entrained in the euphotic zone.[1][2] This creates a high nutrient and high light environment that allows rapid phytoplankton growth.[1][2][7]

This rapid increase in phytoplankton growth occurs because phytoplankton can reproduce rapidly under optimal growth conditions (i.e., high nutrient levels, ideal light and temperature, and minimal losses from grazing and vertical mixing). Eutrophic waters are ideal for spring blooms because their increased nutrient concentration supplies the large quantities of nutrients necessary to promote phytoplankton growth and reproduction.[8] The most limiting nutrient in the marine environment is typically nitrogen (N). This is because most organisms are unable to fix atmospheric nitrogen into usable forms (i.e. ammonium, nitrite, or nitrate). Phosphorus can also be limiting, particularly in freshwater environments and tropical coastal regions.[2] However, it has been shown that in increase in only one of these nutrients does not guarantee an increase in phytoplankton growth. Increasing both nutrients will more reliably lead to an increase in phytoplankton numbers.[8] It can also be argued, with the exception of coastal waters that iron (Fe) is the most limiting nutrient because it is required to fix nitrogen, but is only available in small quantities in the marine environment, coming from dust storms and leaching from rocks.[2] Eutrophic waters have these limiting nutrients in abundance, which allows the phytoplankton in a spring bloom to proliferate rapidly. In terms of temperature, higher growth rates are seen with warmer water temperatures.[9]

Along with thermal stratification, spring blooms can be triggered by salinity stratification due to freshwater input, from sources such as high river runoff. This type of stratification is normally limited to coastal areas and estuaries, including Chesapeake Bay.[10] Freshwater influences primary productivity in two ways. First, because freshwater is less dense, it rests on top of seawater and creates a stratified water column.[1] Second, freshwater often carries nutrients [3] that phytoplankton need to carry out processes, including photosynthesis.

In terms of reproduction, many species of phytoplankton can double at least once per day, allowing for exponential increases in phytoplankton stock size. For example, the stock size of a population that doubles once per day will increase 1000-fold in just 10 days.[2] In addition, there is a lag in the grazing response of herbivorous zooplankton at the start of blooms, which minimize phytoplankton losses. This lag occurs because there is low winter zooplankton abundance and many zooplankton, such as copepods, have longer generation times than phytoplankton.[2]

Spring blooms typically last until late spring or early summer, at which time the bloom collapses due to nutrient depletion in the stratified water column and increased grazing pressure by zooplankton.[1][2][3][5] During winter, wind-driven turbulence and cooling water temperatures break down the stratified water column that formed during the summer. This breakdown allows vertical mixing of the water column and replenishes nutrients from deep water to the surface waters and the rest of the euphotic zone. However, vertical mixing also causes high losses, as phytoplankton are carried below the euphotic zone (so their respiration exceeds primary production). In addition, reduced illumination (intensity and daily duration) during winter limits growth rates.

Alternative mechanisms edit

Historically, blooms have been explained by Sverdrup's critical depth hypothesis, which says blooms are caused by shoaling of the mixed layer. Similarly, Winder and Cloern (2010) described spring blooms as a response to increasing temperature and light availability.[3] However, new explanations have been offered recently, including that blooms occur due to: coupling between phytoplankton growth and zooplankton grazing,[11] the onset of near surface stratification in the spring,[12] mixing of the water column, rather than stratification,[13] and low turbulence.[14]

Coupling between phytoplankton growth and zooplankton grazing edit

A hypothesis that challenges Sverdrup's critical depth hypothesis is the Dilution-Recoupling Hypothesis.[15] This hypothesis argues that stratification is not necessarily required, and instead a deep mixing layer during the winter is more important.[15] This mixing spreads out the phytoplankton and the zooplankton populations, making it less likely for the two to interact. This relaxes the top-down regulation that zooplankton have on phytoplankton. Less interaction leads to less grazing from the zooplankton which then allows the phytoplankton to accumulate a net growth before the spring stratification.[15] Often times the main factor determining the size and duration of the spring bloom is based on the coupling between phytoplankton growth and zooplankton grazing. The bloom will continue to grow as long as phytoplankton growth exceeds zooplankton grazing.[15] The spring bloom will start to decline and then ultimately disappear once zooplankton grazing out scales phytoplankton growth.[15]

Onset of near surface stratification in the spring edit

The stratified layer plays an important role in the phytoplankton distribution of the lake.[16] The depth of the stratified layer determines how far down the phytoplankton can live. This near-surface stratification works together with winter mixing to facilitate phytoplankton growth.[16] Shallow to intermediate stratification after a period of winter mixing produced the largest rates of phytoplankton growth. This is because the stratification helps keep the phytoplankton in areas that receive enough light and nutrients to support growth

Mixing of the water column edit

In lakes, mixing of the water column has shown to have positive effects on phytoplankton growth, and therefore the magnitude of the spring bloom.[16] The greater the magnitude and later onset of the mixing event in winter increases the effect it has. The mixing of the water column is tied into the Dilution-Recoupling hypothesis discussed above that helps to limit zooplankton grazing on phytoplankton. In addition, nutrients such as nitrate in the bottom waters of the lake are mixed with the waters in the top layers. This allows nutrients to be replenished in the upper layer of the water column.[16] These nutrients are crucial for phytoplankton growth. As climate change is predicted to increase temperatures, there is concern that the winter mixing event will not be as strong as it is typically. This could weaken the magnitude of the spring bloom and then negatively impact the food web in lake ecosystems.[16] The decrease in phytoplankton would limit available food for higher trophic levels.

Low Turbulence edit

It has been documented that low surface turbulence due to wind or storms is important for the growth of spring blooms.[17] Related to this is the hypothesis of critical turbulence. This hypothesis says that phytoplankton growth will increase and the spring bloom will begin once the turbulence in the mixed layer weakens. [18] The decrease in turbulence is caused at the end of the winter when air-sea heat fluxes undergo positive changes.[18]

Northward progression edit

At greater latitudes, spring blooms take place later in the year. This northward progression is because spring occurs later, delaying thermal stratification and increases in illumination that promote blooms. A study by Wolf and Woods (1988) showed evidence that spring blooms follow the northward migration of the 12 °C isotherm, suggesting that blooms may be controlled by temperature limitations, in addition to stratification.[1]

At high latitudes, the shorter warm season commonly results in one mid-summer bloom. These blooms tend to be more intense than spring blooms of temperate areas because there is a longer duration of daylight for photosynthesis to take place. Also, grazing pressure tends to be lower because the generally cooler temperatures at higher latitudes slow zooplankton metabolism.[1]

Species succession edit

The spring bloom often consists of a series of sequential blooms of different phytoplankton species. Succession occurs because different species have optimal nutrient uptake at different ambient concentrations and reach their growth peaks at different times. Shifts in the dominant phytoplankton species are likely caused by biological and physical (i.e. environmental) factors.[2] For instance, diatom growth rate becomes limited when the supply of silicate is depleted.[1][2][19] Since silicate is not required by other phytoplankton, such as dinoflagellates, their growth rates continue to increase.

For example, in oceanic environments, diatoms (cells diameter greater than 10 to 70 µm or larger) typically dominate first because they are capable of growing faster. Once silicate is depleted in the environment, diatoms are succeeded by smaller dinoflagellates.[1][2][19] This scenario has been observed in Rhode Island,[20][21][22] as well as Massachusetts and Cape Cod Bay.[7] By the end of a spring bloom, when most nutrients have been depleted, the majority of the total phytoplankton biomass is very small phytoplankton, known as nanoplankton (cell diameter <5 to 10 µm).[2] Nanoplankton can sustain low, but constant stocks, in nutrient depleted environments because they have a larger surface area to volume ratio, which offers a much more effective rate of diffusion.[1][2] The types of phytoplankton comprising a bloom can be determined by examination of the varying photosynthetic pigments found in chloroplasts of each species.[2]

Variability and the influence of climate change edit

Variability in the patterns (e.g., timing of onset, duration, magnitude, position, and spatial extent) of annual spring bloom events has been well documented.[23][24] These variations occur due to fluctuations in environmental conditions, such as wind intensity, temperature, freshwater input, and light. Consequently, spring bloom patterns are likely sensitive to global climate change.[25]

Links have been found between temperature and spring bloom patterns. For example, several studies have reported a correlation between earlier spring bloom onset and temperature increases over time.[23] Furthermore, in Long Island Sound and the Gulf of Maine, blooms begin later in the year, are more productive, and last longer during colder years, while years that are warmer exhibit earlier, shorter blooms of greater magnitude.[24]

Temperature may also regulate bloom sizes. In Narragansett Bay, Rhode Island, a study by Durbin et al. (1992)[26] indicated that a 2 °C increase in water temperature resulted in a three-week shift in the maturation of the copepod, Acartia hudsonica, which could significantly increase zooplankton grazing intensity. Oviatt et al. (2002)[27] noted a reduction in spring bloom intensity and duration in years when winter water temperatures were warmer. Oviatt et al. suggested that the reduction was due to increased grazing pressure, which could potentially become intense enough to prevent spring blooms from occurring altogether. This increase in grazing is dependent on the size of the phytoplankton in the bloom.[25] Copepods are selective feeders based on size and will avoid phytoplankton outside of their size range.[25] This means that grazing pressures from copepods only reduce spring bloom size when the size of the phytoplankton are in their range. This is common in temperate and boreal areas. Warming temperatures have also directly caused the amount of picoplankton (cell diameter <2 µm) to increase while lowering the overall biomass of phytoplankton.[25] These changes are resulting in a shift towards a smaller average phytoplankton size. If this continues the grazing pressure from copepods may decrease due to the phytoplankton being outside of their size range.

There is also evidence of temperature impacting spring blooms in lakes. In Lake Erken, Sweden, it was documented that winter temperatures and duration of ice cover affected the species makeup of the spring bloom.[28] Ice cover impacts the species composition by affecting the amount of light available. In colder winters the spring bloom was dominated by the presence of small centric diatoms, while in warmer winters meroplanktonic diatoms were dominant.[28]

Miller and Harding (2007)[29] suggested climate change (influencing winter weather patterns and freshwater influxes) was responsible for shifts in spring bloom patterns in the Chesapeake Bay. They found that during warm, wet years (as opposed to cool, dry years), the spatial extent of blooms was larger and was positioned more seaward. Also, during these same years, biomass was higher and peak biomass occurred later in the spring.

Environmental Effects edit

If a spring bloom occurs in an area with well established and rigid stratification it can become a harmful algal bloom.[30] When the phytoplankton die, they sink to lower areas in the water column. There they are decomposed by bacteria that consume all of the oxygen at that depth through respiration. Since the water is stratified, there is no addition of new oxygen to these depths.[30] This area is called a dead zone and the lack of oxygen can lead to the deaths of many aquatic organisms. Some harmful algal blooms release toxins that make the water unsafe for drinking, seafood unsafe for eating, and conditions even more unfavorable for aquatic species to live.[31]

 
A harmful algal bloom in Lake Erie during 2011







Spring Blooms can also act as a carbon sink.[15] Through the process of photosynthesis phytoplankton uptake carbon dioxide. This removes carbon out of the atmosphere and stores it as biomass. When the phytoplankton die, they sink to the bottom of the water column where carbon becomes trapped.[15]

See also[edit] edit

  1. ^ a b c d e f g h i j k Mann, K.H., Lazier, J.R.N. (2006). Dynamics of Marine Ecosystems: Biological-Physical Interactions in the Oceans. Oxford: Blackwell Publishing Ltd. ISBN 1-4051-1118-6
  2. ^ a b c d e f g h i j k l m n o Miller, C.B. (2004). "Biological Oceanography" Oxford: Blackwell Publishing Ltd. ISBN 978-0-632-05536-4
  3. ^ a b c d Winder, M. and Cloern, J.E. (2010). "The annual cycles of phytoplankton biomass". Philosophical Transactions of the Royal Society B 365: 3215–3226. doi:10.1098/rstb.2010.0125
  4. ^ Oviatt, C., Keller, A., and Reed, L. (2002). "Annual Primary Production in Narragansett Bay with no Bay-Wide Winter–Spring Phytoplankton Bloom". Estuarine, Coastal and Shelf Science 54: 1013–1026. doi:10.1006/ecss.2001.0872
  5. ^ a b Smayda, T.J. (1998). "Patterns of variability characterizing marine phytoplankton, with examples from Narragansett Bay". ICES Journal of Marine Science 55: 562–573
  6. ^ Hunter-Cevera, Kristen R.; Neubert, Michael G.; Olson, Robert J.; Solow, Andrew R.; Shalapyonok, Alexi; Sosik, Heidi M. (2016). "Physiological and ecological drivers of early spring blooms of a coastal phytoplankter". Science. 354 (6310): 326–329. Bibcode:2016Sci...354..326H. doi:10.1126/science.aaf8536. PMID 27846565.
  7. ^ a b Hunt, C.D., Borkman, D.G., Libby, P.S., Lacouture, R., Turner, J.T., and Mickelson, M.J. (2010). "Phytoplankton Patterns in Massachusetts Bay—1992–2007". Estuaries and Coasts 33: 448–470. doi:10.1007/s12237-008-9125-9
  8. ^ a b Su, Xiaomei; Steinman, Alan D.; Oudsema, Maggie; Hassett, Michael; Xie, Liqiang (2019). "The influence of nutrients limitation on phytoplankton growth and microcystins production in Spring Lake, USA". Chemosphere. 234: 34–42. doi:10.1016/j.chemosphere.2019.06.047.
  9. ^ Deng, Jianming; Qin, Boqiang; Paerl, Hans W.; Zhang, Yunlin; Wu, Pan; Ma, Jianrong; Chen, Yuwei (2014-12-02). "Effects of Nutrients, Temperature and Their Interactions on Spring Phytoplankton Community Succession in Lake Taihu, China". PLOS ONE. 9 (12): e113960. doi:10.1371/journal.pone.0113960. ISSN 1932-6203. PMC 4252073. PMID 25464517.{{cite journal}}: CS1 maint: PMC format (link) CS1 maint: unflagged free DOI (link)
  10. ^ Harding, L. W. and Perry, E. S. (1997). "Long-term increase of phytoplankton biomass in Chesapeake Bay, 1950–94." Marine Ecological Progress Series 157: 39–52. doi:10.3354/meps157039
  11. ^ Behrenfeld, M.J. (2010). "Abandoning Sverdrup's Critical Depth Hypothesis on phytoplankton blooms". Ecology 91:977–989. doi:10.1890/09-1207.1
  12. ^ Chiswell, S. M., 2011, "The spring phytoplankton bloom: don’t abandon Sverdrup completely": Marine Ecology Progress Series, v. 443, p. 39–50 – doi:10.3354/meps09453
  13. ^ Townsend, D.W., Cammen, L.M., Holligan, P.M., Campbell, D.E., Pettigrew, N.R. (1994). "Causes and consequences of variability in the timing of spring phytoplankton blooms". Deep-Sea Research 41: 747–765
  14. ^ Huisman, J., van Oostveen, P., Weissing, F.J. (1999). "Critical depth and critical turbulence: two different mechanisms for the development of phytoplankton blooms." Limnological Oceanography 44: 1781–1787
  15. ^ a b c d e f g George, Jennifer A.; Lonsdale, Darcy J.; Merlo, Lucas R.; Gobler, Christopher J. (2015). "The interactive roles of temperature, nutrients, and zooplankton grazing in controlling the winter-spring phytoplankton bloom in a temperate, coastal ecosystem, Long Island Sound: Long Island Sound winter-spring bloom". Limnology and Oceanography. 60 (1): 110–126. doi:10.1002/lno.10020.
  16. ^ a b c d e González-Gil, Ricardo; González Taboada, Fernando; Cáceres, Carlos; Largier, John L.; Anadón, Ricardo (2018). "Winter-mixing preconditioning of the spring phytoplankton bloom in the Bay of Biscay: Winter mixing and spring bloom". Limnology and Oceanography. 63 (3): 1264–1282. doi:10.1002/lno.10769.
  17. ^ Garcia-Soto, Carlos; Pingree, Robin D. (2009). "Spring and summer blooms of phytoplankton (SeaWiFS/MODIS) along a ferry line in the Bay of Biscay and western English Channel". Continental Shelf Research. 29 (8): 1111–1122. doi:10.1016/j.csr.2008.12.012.
  18. ^ a b Carreto, José I.; Montoya, Nora G.; Carignan, Mario O.; Akselman, Rut; Acha, E. Marcelo; Derisio, Carla (2016). "Environmental and biological factors controlling the spring phytoplankton bloom at the Patagonian shelf-break front – Degraded fucoxanthin pigments and the importance of microzooplankton grazing". Progress in Oceanography. 146: 1–21. doi:10.1016/j.pocean.2016.05.002.
  19. ^ a b Kristiansen, S., Farbrot, T., and Naustvoll, L. (2001). "Spring bloom nutrient dynamics in the Oslofjord". Marine Ecology Progress Series 219: 41–49
  20. ^ Smayda, T.J.(1957). "Phytoplankton studies in lower Narragansett Bay". Limnology and Oceanography 2(4) 342-359
  21. ^ Nixon, S.W., Fulweiler, R.W., Buckley, B.A., Granger, S.L., Nowicki, B.L., Henry, K.M. (2009). "The impact of changing climate on phenology, productivity, and benthic-pelagic coupling in Narragansett Bay". Estuarine, Coastal and Shelf Science 82: 1-18
  22. ^ Pratt, D.M.(1959). "The phytoplankton of Narragansett Bay". Limnology and Oceanography 4(4) 425-440
  23. ^ a b Winder, M. and Cloern, J.E. (2010). "The annual cycles of phytoplankton biomass". Philosophical Transactions of the Royal Society B 365: 3215–3226. doi:10.1098/rstb.2010.0125
  24. ^ a b Smayda, T.J. (1998). "Patterns of variability characterizing marine phytoplankton, with examples from Narragansett Bay". ICES Journal of Marine Science 55: 562–573
  25. ^ a b c d Sommer, Ulrich; Aberle, Nicole; Lengfellner, Kathrin; Lewandowska, Aleksandra (2012). "The Baltic Sea spring phytoplankton bloom in a changing climate: an experimental approach". Marine Biology. 159 (11): 2479–2490. doi:10.1007/s00227-012-1897-6. ISSN 0025-3162.
  26. ^ Durbin, A.G. and Durbin, E.G. (1992). "Seasonal changes in size frequency distribution and estimated age in the marine copepod Acartia hudsortica during a winter-spring diatom bloom in Narragansett Bay". Limnol. Oceanogr., 37(2): 379–392
  27. ^ Oviatt, C., Keller, A., and Reed, L. (2002). "Annual Primary Production in Narragansett Bay with no Bay-Wide Winter–Spring Phytoplankton Bloom". Estuarine, Coastal and Shelf Science 54: 1013–1026. doi:10.1006/ecss.2001.0872
  28. ^ a b Yang, Yang; Stenger-Kovács, Csilla; Padisák, Judit; Pettersson, Kurt (2016). "Effects of winter severity on spring phytoplankton development in a temperate lake (Lake Erken, Sweden)". Hydrobiologia. 780 (1): 47–57. doi:10.1007/s10750-016-2777-8. ISSN 0018-8158.
  29. ^ Miller, W.D. and Harding Jr., L.W. (2007). "Climate forcing of the spring bloom in Chesapeake Bay". Marine Ecology Progress Series 331: 11–22
  30. ^ a b Joyce, S (2000). "The dead zones: oxygen-starved coastal waters". Environmental Health Perspectives. 108 (3). doi:10.1289/ehp.108-a120. ISSN 0091-6765.
  31. ^ Pal, Mili; Yesankar, Prerna J.; Dwivedi, Ajay; Qureshi, Asifa (2020). "Biotic control of harmful algal blooms (HABs): A brief review". Journal of Environmental Management. 268: 110687. doi:10.1016/j.jenvman.2020.110687.