Marine protists

Marine protists are defined by their habitat as protists that live in marine environments, that is, in the saltwater of seas or oceans or the brackish water of coastal estuaries. Life originated as single-celled prokaryotes (bacteria and archaea) and later evolved into more complex eukaryotes. Eukaryotes are the more developed life forms known as plants, animals, fungi and protists. Protists are the eukaryotes that cannot be classified as plants, fungi or animals. They are usually single-celled and microscopic. The term protist came into use historically as a term of convenience for eukaryotes that cannot be strictly classified as plants, animals or fungi. They are not a part of modern cladistics, because they are paraphyletic (lacking a common ancestor).

Marine protists
Alga (phytoplankton)
Protozoan (zooplankton)
Protists are usually one-celled microorganisms. They include algae (autotrophs which make their own food) and protozoans (heterotrophs which eat the algae for food). But many protists are in between (mixotrophs which are a bit of both).

Most protists are too small to be seen with the naked eye. They are highly diverse organisms currently organised into 18 phyla, but not easy to classify.[1][2] Studies have shown high protist diversity exists in oceans, deep sea-vents and river sediments, suggesting large numbers of eukaryotic microbial communities have yet to be discovered.[3][4] There has been little research on mixotrophic protists, but recent studies in marine environments found mixotrophic protests contribute a significant part of the protist biomass.[5] Since protists are eukaryotes (and not prokaryotes) they possess within their cell at least one nucleus, as well as organelles such as mitochondria and Golgi bodies. Protists are asexual but can reproduce rapidly through mitosis or by fragmentation.

In contrast to the cells of prokaryotes, the cells of eukaryotes are highly organised. Plants, animals and fungi are usually multi-celled and are typically macroscopic. Most protists are single-celled and microscopic. But there are exceptions. Some single-celled marine protists are macroscopic. Some marine slime molds have unique life cycles that involve switching between unicellular, colonial, and multicellular forms.[6] Other marine protist are neither single-celled nor microscopic, such as seaweed.

Protists have been described as a taxonomic grab bag of misfits where anything that doesn't fit into one of the main biological kingdoms can be placed.[7] Some modern authors prefer to exclude multicellular organisms from the traditional definition of a protist, restricting protists to unicellular organisms.[8][9] This more constrained definition excludes many brown, multicellular red and green algae, and slime molds.[10]

Background

Phylogenetic and symbiogenetic tree of living organisms, showing a schematic view of the central position occupied by the protista (protists)

Trophic modes

Protists can be broadly divided into four groups depending on whether their nutrition is plant-like, animal-like, fungal-like,[11] or a mixture of these.[12]

Protists according to how they get food
Type of protist Description Example Some other examples
Plant-like Autotrophic protists that make their own food without needing to consume other organisms, usually by photosynthesis (sometimes by chemosynthesis) Green algae, Pyramimonas Red and brown algae, diatoms, coccolithophores and some dinoflagellates. Plant-like protists are important components of phytoplankton discussed below.
Animal-like Heterotrophic protists that get their food consuming other organisms (bacteria, archaea and small algae) Radiolarian protist as drawn by Haeckel Foraminiferans, and some marine amoebae, ciliates and flagellates.
Fungal-like Saprotrophic protists that get their food from the remains of organisms that have broken down and decayed Marine slime nets form labyrinthine networks of tubes in which amoeba without pseudopods can travel Marine lichen
Mixotrophs
Various
(see below)
Mixotrophic and osmotrophic protists that get their food from a combination of the above Euglena mutabilis, a photosynthetic flagellate Many marine mixotrops are found among protists, particularly among ciliates and dinoflagellates[5]
micrograph
cell schematic
Choanoflagellates, unicellular "collared" flagellate protists, are thought to be the closest living relatives of the animals.[13]
External video
How microscopic hunters get their lunch
Euglenoids: Single-celled shapeshifters
How do protozoans get around?

The fungus-like protist saprobes are specialized to absorb nutrients from nonliving organic matter, such as dead organisms or their wastes. For instance, many types of oomycetes grow on dead animals or algae. Marine saprobic protists have the essential function of returning inorganic nutrients to the water. This process allows for new algal growth, which in turn generates sustenance for other organisms along the food chain. Indeed, without saprobe species, such as protists, fungi, and bacteria, life would cease to exist as all organic carbon became "tied up" in dead organisms.[15][16]

Mixotrophs

Mixotrophic radiolarians
Acantharian radiolarian hosts Phaeocystis symbionts
White Phaeocystis algal foam washing up on a beach

Mixotrophs have no single trophic mode. A mixotroph is an organism that can use a mix of different sources of energy and carbon, instead of having a single trophic mode on the continuum from complete autotrophy at one end to heterotrophy at the other. It is estimated that mixotrophs comprise more than half of all microscopic plankton.[17] There are two types of eukaryotic mixotrophs: those with their own chloroplasts, and those with endosymbionts—and others that acquire them through kleptoplasty or by enslaving the entire phototrophic cell.[18]

The distinction between plants and animals often breaks down in very small organisms. Possible combinations are photo- and chemotrophy, litho- and organotrophy, auto- and heterotrophy or other combinations of these. Mixotrophs can be either eukaryotic or prokaryotic.[19] They can take advantage of different environmental conditions.[20]

Recent studies of marine microzooplankton found 30–45% of the ciliate abundance was mixotrophic, and up to 65% of the amoeboid, foram and radiolarian biomass was mixotrophic.[5]

Phaeocystis is an important algal genus found as part of the marine phytoplankton around the world. It has a polymorphic life cycle, ranging from free-living cells to large colonies.[21] It has the ability to form floating colonies, where hundreds of cells are embedded in a gel matrix, which can increase massively in size during blooms.[22] As a result, Phaeocystis is an important contributor to the marine carbon[23] and sulfur cycles.[24] Phaeocystis species are endosymbionts to acantharian radiolarians.[25][26]

Mixotrophic plankton that combine phototrophy and heterotrophy – table based on Stoecker et. al., 2017[27]
General types Description Example Further examples
Bacterioplankton Photoheterotrophic bacterioplankton Vibrio cholerae Roseobacter spp.
Erythrobacter spp.
Gammaproteobacterial clade OM60
Widespread among bacteria and archaea
Phytoplankton Called constitutive mixotrophs by Mitra et. al., 2016.[28] Phytoplankton that eat: photosynthetic protists with inherited plastids and the capacity to ingest prey. Ochromonas species Ochromonas spp.
Prymnesium parvum
Dinoflagellate examples: Fragilidium subglobosum,Heterocapsa triquetra,Karlodinium veneficum,Neoceratium furca,Prorocentrum minimum
Zooplankton Called nonconstitutive mixotrophs by Mitra et. al., 2016.[28] Zooplankton that are photosynthetic: microzooplankton or metazoan zooplankton that acquire phototrophy through chloroplast retentiona or maintenance of algal endosymbionts.
Generalists Protists that retain chloroplasts and rarely other organelles from many algal taxa Most oligotrich ciliates that retain plastidsa
Specialists 1. Protists that retain chloroplasts and sometimes other organelles from one algal species or very closely related algal species Dinophysis acuminata Dinophysis spp.
Mesodinium rubrum
2. Protists or zooplankton with algal endosymbionts of only one algal species or very closely related algal species Noctiluca scintillans Metazooplankton with algal endosymbionts
Most mixotrophic Rhizaria (Acantharea, Polycystinea, and Foraminifera)
Green Noctiluca scintillans
aChloroplast (or plastid) retention = sequestration = enslavement. Some plastid-retaining species also retain other organelles and prey cytoplasm.

Protist locomotion

Another way of categorising protists is according to their mode of locomotion. Many unicellular protists, particularly protozoans, are motile and can generate movement using flagella, cilia or pseudopods. Cells which use flagella for movement are usually referred to as flagellates, cells which use cilia are usually referred to as ciliates, and cells which use pseudopods are usually referred to as amoeba or amoeboids. Other protists are not motile, and consequently have no movement mechanism.

Protists according to how they move
Type of protist Movement mechanism Description Example Other examples
Motile Flagellates A flagellum (Latin for whip) is a lash-like appendage that protrudes from the cell body of some protists (as well as some bacteria). Flagellates use from one to several flagella for locomotion and sometimes as feeding and sensory organelle. Cryptophytes All dinoflagellates and nanoflagellates (choanoflagellates, silicoflagellates, most green algae)[29][30]
(Other protists go through a phase as gametes when they have temporary flagellum – some radiolarians, foraminiferans and Apicomplexa)
Ciliates A cilium (Latin for eyelash) is a tiny flagellum. Ciliates use multiple cilia, which can number in many hundreds, to power themselves through the water. Paramecium bursaria
click to see cilia
Foraminiferans, and some marine amoebae, ciliates and flagellates.
Amoebas
(amoeboids)
Pseudopods (Greek for false feet) are lobe-like appendages which amoebas use to anchor to a solid surface and pull themselves forward. They can change their shape by extending and retracting these pseudopods.[31]
Amoeba Found in every major protist lineage. Amoeboid cells occur among the protozoans, but also in the algae and the fungi.[32][33]
Not motile
none
Diatom Diatoms, coccolithophores, and non‐motile species of Phaeocystis[30] Among protozoans the parasitic Apicomplexa are non‐motile.
Difference of beating pattern of flagellum and cilium

Flagella are used in prokaryotes (archaea and bacteria) as well as protists. In addition, both flagella and cilia are widely used in eukaryotic cells (plant and animal) apart from protists.

The regular beat patterns of eukaryotic cilia and flagella generates motion on a cellular level. Examples range from the propulsion of single cells such as the swimming of spermatozoa to the transport of fluid along a stationary layer of cells such as in a respiratory tract. Though eukaryotic flagella and motile cilia are ultrastructurally identical, the beating pattern of the two organelles can be different. In the case of flagella, the motion is often planar and wave-like, whereas the motile cilia often perform a more complicated three-dimensional motion with a power and recovery stroke.

Eukaryotic flagella—those of animal, plant, and protist cells—are complex cellular projections that lash back and forth. Eukaryotic flagella are classed along with eukaryotic motile cilia as undulipodia[34] to emphasize their distinctive wavy appendage role in cellular function or motility. Primary cilia are immotile, and are not undulipodia.

Marine flagellates from the genera (left to right)
Cryptaulax, Abollifer, Bodo, Rhynchomonas, Kittoksia, Allas, and Metromonas[35]
Cilia performs powerful forward strokes with a stiffened flagellum followed by relatively slow recovery movement with a relaxed flagellum

Ciliates generally have hundreds to thousands of cilia that are densely packed together in arrays. Like the flagella, the cilia are powered by specialised molecular motors. An efficient forward stroke is made with a stiffened flagellum, followed by an inefficient backward stroke made with a relaxed flagellum. During movement, an individual cilium deforms as it uses the high-friction power strokes and the low-friction recovery strokes. Since there are multiple cilia packed together on an individual organism, they display collective behaviour in a metachronal rhythm. This means the deformation of one cilium is in phase with the deformation of its neighbor, causing deformation waves that propagate along the surface of the organism. These propagating waves of cilia are what allow the organism to use the cilia in a coordinated manner to move. A typical example of a ciliated microorganism is the Paramecium, a one-celled, ciliated protozoan covered by thousands of cilia. The cilia beating together allow the Paramecium to propel through the water at speeds of 500 micrometers per second.[36]

External video
Paramecium: The White Rat of Ciliates

Marine algae

Algae is an informal term for a widespread and diverse group of photosynthetic protists which are not necessarily closely related and are thus polyphyletic. Marine algae can be divided into six groups: green, red and brown algae, euglenophytes, dinoflagellates and diatoms.

Dinoflagellates and diatoms are important components of marine algae and have their own sections below. Euglenophytes are a phylum of unicellular flagellates with only a few marine members.

Not all algae are microscopic. Green, red and brown algae all have multicellular macroscopic forms that make up the familiar seaweeds. Green algae, an informal group, contains about 8,000 recognised species.[37] Many species live most of their lives as single cells or are filamentous, while others form colonies made up from long chains of cells, or are highly differentiated macroscopic seaweeds. Red algae, a (disputed) phylum contains about 7,000 recognised species,[38] mostly multicellular and including many notable seaweeds.[38][39] Brown algae form a class containing about 2,000 recognised species,[40] mostly multicellular and including many seaweeds such as kelp. Unlike higher plants, algae lack roots, stems, or leaves. They can be classified by size as microalgae or macroalgae.

Microalgae are the microscopic types of algae, not visible to the naked eye. They are mostly unicellular species which exist as individuals or in chains or groups, though some are multicellular. Microalgae are important components of the marine protists discussed above, as well as the phytoplankton discussed below. They are very diverse. It has been estimated there are 200,000-800,000 species of which about 50,000 species have been described.[41] Depending on the species, their sizes range from a few micrometers (µm) to a few hundred micrometers. They are specially adapted to an environment dominated by viscous forces.

Macroalgae are the larger, multicellular and more visible types of algae, commonly called seaweeds. Seaweeds usually grow in shallow coastal waters where they are anchored to the seafloor by a holdfast. Like microalgae, macroalgae (seaweeds) can be regarded as marine protists since they are not true plants. But they are not microorganisms, so they are not within the scope of this article.

Unicellular organisms are usually microscopic, less than one tenth of a millimeter long. There are exceptions. Mermaid's wineglass, a genus of subtropical green algae, is single-celled but remarkably large and complex in form with a single large nucleus, making it a model organism for studying cell biology.[43] Another single-celled algae, Caulerpa taxifolia, has the appearance of a vascular plant including "leaves" arranged neatly up stalks like a fern. Selective breeding in aquariums to produce hardier strains resulted in an accidental release into the Mediterranean where it has become an invasive species known colloquially as killer algae.[44]

Diatoms

Diatoms come in many shapes

Diatoms are photosynthetic unicellular algae populating the oceans and other waters around the globe. They form a (disputed) phylum containing about 100,000 recognised species. Diatoms generate about 20 percent of all oxygen produced on the planet each year,[14] and take in over 6.7 billion metric tons of silicon each year from the waters in which they live.[45] They produce 25–45% of the total primary production of organic material in the oceans,[46][47][48] owing to their prevalence in open-ocean regions when total phytoplankton biomass is maximal.[49][50]

Diatoms are enclosed in protective silica (glass) shells called frustules. They are classified by the shape of these glass cages in which they live, and which they build as they grow. Each frustule is made from two interlocking parts covered with tiny holes through which the diatom exchanges nutrients and wastes.[51] Dead diatoms drift to the ocean floor where, over millions of years, the remains of their frustules can build up as much as half a mile deep.[52] Diatoms have relatively high sinking speeds compared with other phytoplankton groups, and they account for about 40% of particulate carbon exported to ocean depths.[48][53][50]

Diatom shapes
          Drawings by Haeckel 1904 (click for details)
Diatoms
Diatoms have a silica shell (frustule) with radial (centric) or bilateral (pennate) symmetry
External video
Diatoms: Tiny factories you can see from space
Diatom 3D interference contrast
Structure of a centric diatom frustule[54]

Physically driven seasonal enrichments in surface nutrients favour diatom blooms. Anthropogenic climate change will directly affect these seasonal cycles, changing the timing of blooms and diminishing their biomass, which will reduce primary production and CO2 uptake.[55][50] Remote sensing data suggests there was a global decline of diatoms between 1998 and 2012, particularly in the North Pacific, associated with shallowing of the surface mixed layer and lower nutrient concentrations.[56][50]

Coccolithophores

Coccolithophores
...have plates called coccoliths
...extinct fossil
Coccolithophores build calcite skeletons important to the marine carbon cycle[59]

Coccolithophores are minute unicellular photosynthetic protists with two flagella for locomotion. Most of them are protected by calcium carbonate shells covered with ornate circular plates or scales called coccoliths. The term coccolithophore derives from the Greek for a seed carrying stone, referring to their small size and the coccolith stones they carry. Under the right conditions they bloom, like other phytoplankton, and can turn the ocean milky white.[60]

The fossil coccolithophore Braarudosphaera bigelowii has an unusual shell with a regular dodecahedral structure about 10 micrometers across.[61]

Dinoflagellates

Dinoflagellate shapes
Unarmored dinoflagellates Kofoid (1921)
Haeckel Peridinea (1904)

Dinoflagellates are usually positioned as part of the algae group, and form a phylum of unicellular flagellates with about 2,000 marine species.[62] The name comes from the Greek "dinos" meaning whirling and the Latin "flagellum" meaning a whip or lash. This refers to the two whip-like attachments (flagella) used for forward movement. Most dinoflagellates are protected with red-brown, cellulose armour. Like other phytoplankton, dinoflagellates are r-strategists which under right conditions can bloom and create red tides. Excavates may be the most basal flagellate lineage.[29]

By trophic orientation dinoflagellates are all over the place. Some dinoflagellates are known to be photosynthetic, but a large fraction of these are in fact mixotrophic, combining photosynthesis with ingestion of prey (phagotrophy).[63] Some species are endosymbionts of marine animals and other protists, and play an important part in the biology of coral reefs. Others predate other protozoa, and a few forms are parasitic. Many dinoflagellates are mixotrophic and could also be classified as phytoplankton.

The toxic dinoflagellate Dinophysis acuta acquire chloroplasts from its prey. "It cannot catch the cryptophytes by itself, and instead relies on ingesting ciliates such as the red Mesodinium rubrum, which sequester their chloroplasts from a specific cryptophyte clade (Geminigera/Plagioselmis/Teleaulax)".[27]

A surf wave at night sparkles with blue light due to the presence of a bioluminescent dinoflagellate, such as Lingulodinium polyedrum
Suggested explanation for glowing seas[64]
Dinoflagellates
        Armoured
        Unarmoured
Traditionally dinoflagellates have been presented as armoured or unarmoured

Dinoflagellates often live in symbiosis with other organisms. Many nassellarian radiolarians house dinoflagellate symbionts within their tests.[65] The nassellarian provides ammonium and carbon dioxide for the dinoflagellate, while the dinoflagellate provides the nassellarian with a mucous membrane useful for hunting and protection against harmful invaders.[66] There is evidence from DNA analysis that dinoflagellate symbiosis with radiolarians evolved independently from other dinoflagellate symbioses, such as with foraminifera.[67]

Some dinoflagellates are bioluminescent. At night, ocean water can light up internally and sparkle with blue light because of these dinoflagellates.[68][69] Bioluminescent dinoflagellates possess scintillons, individual cytoplasmic bodies which contain dinoflagellate luciferase, the main enzyme involved in the luminescence. The luminescence, sometimes called the phosphorescence of the sea, occurs as brief (0.1 sec) blue flashes or sparks when individual scintillons are stimulated, usually by mechanical disturbances from, for example, a boat or a swimmer or surf.[70]

Marine protozoans

Protozoans are protists which feed on organic matter such as other microorganisms or organic tissues and debris.[74][75] Historically, the protozoa were regarded as "one-celled animals", because they often possess animal-like behaviours, such as motility and predation, and lack a cell wall, as found in plants and many algae.[76][77] Although the traditional practice of grouping protozoa with animals is no longer considered valid, the term continues to be used in a loose way to identify single-celled organisms that can move independently and feed by heterotrophy.

Marine protozoans include zooflagellates, foraminiferans, radiolarians and some dinoflagellates.

Radiolarians

Radiolarian shapes
          Drawings by Haeckel 1904 (click for details)

Radiolarians are unicellular predatory protists encased in elaborate globular shells, typically between 0.1 and 0.2 millimetres in size, usually made of silica and pierced with holes. Their name comes from the Latin for "radius". They catch prey by extending parts of their body through the holes. As with the silica frustules of diatoms, radiolarian shells can sink to the ocean floor when radiolarians die and become preserved as part of the ocean sediment. These remains, as microfossils, provide valuable information about past oceanic conditions.[78]

Turing and radiolarian morphology
Shell of a spherical radiolarian
Shell micrographs
Computer simulations of Turing patterns on a sphere
closely replicate some radiolarian shell patterns[79]
External video
Radiolarian geometry
Ernst Haeckel's radiolarian engravings

Foraminiferans

Foraminiferan shapes
          Drawings by Haeckel 1904 (click for details)

Like radiolarians, foraminiferans (forams for short) are single-celled predatory protists, also protected with shells that have holes in them. Their name comes from the Latin for "hole bearers". Their shells, often called tests, are chambered (forams add more chambers as they grow). The shells are usually made of calcite, but are sometimes made of agglutinated sediment particles or chiton, and (rarely) of silica. Most forams are benthic, but about 40 species are planktic.[80] They are widely researched with well established fossil records which allow scientists to infer a lot about past environments and climates.[78]

Foraminiferans
...can have more than one nucleus
...and defensive spines
Foraminiferans are important unicellular zooplankton protists, with calcium tests
External video
foraminiferans
Foraminiferal networks and growth

A number of forams are mixotrophic (see below). These have unicellular algae as endosymbionts, from diverse lineages such as the green algae, red algae, golden algae, diatoms, and dinoflagellates.[80] Mixotrophic foraminifers are particularly common in nutrient-poor oceanic waters.[82] Some forams are kleptoplastic, retaining chloroplasts from ingested algae to conduct photosynthesis.[83]

Amoeba

Shelled and naked amoeba
                  Amoeba can be shelled (testate) or naked
External video
Amoebas
Testate amoebas
Feeding amoebas
Ciliate shapes
          Drawings by Haeckel 1904 (click for details)

Ciliates

Conjugation of two Coleps sp.
Two similar-looking but sexually distinct partners connected at their front ends exchange genetic material via a plasma bridge.
External video
Peritrich Ciliates

Macroscopic protists

Protist shells

Diatoms
Diatoms, major components of marine plankton, have silica skeletons called frustules. "The microscopic structures of diatoms help them manipulate light, leading to hopes they could be used in new technologies for light detection, computing or robotics.[86]
SEM images of pores in diatom frustules[87]

Many protists have protective shells, usually made from calcium carbonate or silica (glass).

Diatom shells are called frustules and are made from silica. These glass structures have accumulated for over 100 million years leaving rich deposits of nano and microstructured silicon oxide in the form of diatomaceous earth around the globe. The evolutionary causes for the generation of nano and microstructured silica by photosynthetic algae are not yet clear. However, in 2018 it was shown that reflection of ultraviolet light by nanostructured silica protects the DNA in the algal cells, and this may be an evolutionary cause for the formation of the glass cages.[87][88]

Fossil radiolarian
X-ray microtomography of Triplococcus acanthicus
This is a microfossil from the Middle Ordovician with four nested spheres. The innermost sphere is highlighted red. Each segment is shown at the same scale.[89]
Benefits of coccolithophore calcification
(A) Accelerated photosynthesis includes CCM (1) and enhanced light uptake via scattering of scarce photons for deep-dwelling species (2). (B) Protection from photodamage includes sunshade protection from ultraviolet (UV) light and photosynthetic active radiation (PAR) (1) and energy dissipation under high-light conditions (2). (C) Armor protection includes protection against viral/bacterial infections (1) and grazing by selective (2) and nonselective (3) grazers.[90]
Triparma laevis and a drawing of its silicate shell, scale bar = 1 μm.
Exploded drawing of the shell, D = dorsal plate, G = girdle plate, S = shield plate and V = ventral plate.
Triparma laevis belongs to the Bolidophyceae, a sister taxon to the diatoms.[91][92]
Relative sizes of coccolithophores
Size comparison between the relatively large coccolithophore Scyphosphaera apsteinii and the relatively small but ubiquitous coccolithophore Emiliania huxleyi[93]
Energetic costs of coccolithophore calcification
Energetic costs are reported in percentage of total photosynthetic budget. (A) Transport processes include the transport into the cell from the surrounding seawater of primary calcification substrates Ca2+ and HCO3− (black arrows) and the removal of the end product H+ from the cell (gray arrow). The transport of Ca2+ through the cytoplasm to the CV is the dominant cost associated with calcification (Table 1). (B) Metabolic processes include the synthesis of CAPs (gray rectangles) by the Golgi complex (white rectangles) that regulate the nucleation and geometry of CaCO3 crystals. The completed coccolith (gray plate) is a complex structure of intricately arranged CAPs and CaCO3 crystals. (C) Mechanical and structural processes account for the secretion of the completed coccoliths that are transported from their original position adjacent to the nucleus to the cell periphery, where they are transferred to the surface of the cell. The costs associated with these processes are likely to be comparable to organic-scale exocytosis in noncalcifying haptophyte algae.[90]
  • Xu, K., Hutchins, D. and Gao, K. (2018) "Coccolith arrangement follows Eulerian mathematics in the coccolithophore Emiliania huxleyi". PeerJ, 6: e4608. doi:10.1126/science.aaa7378.

References

  1. Cavalier-Smith T (December 1993). "Kingdom protozoa and its 18 phyla". Microbiological Reviews. 57 (4): 953–94. doi:10.1128/mmbr.57.4.953-994.1993. PMC 372943. PMID 8302218.
  2. Corliss JO (1992). "Should there be a separate code of nomenclature for the protists?". BioSystems. 28 (1–3): 1–14. doi:10.1016/0303-2647(92)90003-H. PMID 1292654.
  3. Slapeta J, Moreira D, López-García P (2005). "The extent of protist diversity: insights from molecular ecology of freshwater eukaryotes". Proceedings of the Royal Society B: Biological Sciences. 272 (1576): 2073–81. doi:10.1098/rspb.2005.3195. PMC 1559898. PMID 16191619.
  4. Moreira D, López-García P (2002). "The molecular ecology of microbial eukaryotes unveils a hidden world" (PDF). Trends in Microbiology. 10 (1): 31–8. doi:10.1016/S0966-842X(01)02257-0. PMID 11755083.
  5. Leles, S.G.; Mitra, A.; Flynn, K.J.; Stoecker, D.K.; Hansen, P.J.; Calbet, A.; McManus, G.B.; Sanders, R.W.; Caron, D.A.; Not, F.; Hallegraeff, G.M. (2017). "Oceanic protists with different forms of acquired phototrophy display contrasting biogeographies and abundance". Proceedings of the Royal Society B: Biological Sciences. 284 (1860): 20170664. doi:10.1098/rspb.2017.0664. PMC 5563798. PMID 28768886.
  6. Devreotes P (1989). "Dictyostelium discoideum: a model system for cell-cell interactions in development". Science. 245 (4922): 1054–8. Bibcode:1989Sci...245.1054D. doi:10.1126/science.2672337. PMID 2672337.
  7. Neil A C, Reece J B, Simon E J (2004) Essential biology with physiology Pearson/Benjamin Cummings, Page 291. ISBN 9780805375039
  8. O'Malley MA, Simpson AG, Roger AJ (2012). "The other eukaryotes in light of evolutionary protistology". Biology & Philosophy. 28 (2): 299–330. doi:10.1007/s10539-012-9354-y. S2CID 85406712.
  9. Adl SM, Simpson AG, Farmer MA, Andersen RA, Anderson OR, Barta JR, Bowser SS, Brugerolle G, Fensome RA, Fredericq S, James TY, Karpov S, Kugrens P, Krug J, Lane CE, Lewis LA, Lodge J, Lynn DH, Mann DG, McCourt RM, Mendoza L, Moestrup O, Mozley-Standridge SE, Nerad TA, Shearer CA, Smirnov AV, Spiegel FW, Taylor MF (2005). "The new higher level classification of eukaryotes with emphasis on the taxonomy of protists". The Journal of Eukaryotic Microbiology. 52 (5): 399–451. doi:10.1111/j.1550-7408.2005.00053.x. PMID 16248873. S2CID 8060916.
  10. Margulis L, Chapman MJ (19 March 2009). Kingdoms and Domains: An Illustrated Guide to the Phyla of Life on Earth. Academic Press. ISBN 9780080920146.
  11. Whittaker, R.H.; Margulis, L. (1978). "Protist classification and the kingdoms of organisms". Biosystems. 10 (1–2): 3–18. doi:10.1016/0303-2647(78)90023-0. PMID 418827.
  12. Faure, E; Not, F; Benoiston, AS; Labadie, K; Bittner, L; Ayata, SD (2019). "Mixotrophic protists display contrasted biogeographies in the global ocean". ISME Journal. 13 (4): 1072–1083. doi:10.1038/s41396-018-0340-5. PMC 6461780. PMID 30643201.
  13. Budd, Graham E; Jensen, Sören (2017). "The origin of the animals and a 'Savannah' hypothesis for early bilaterian evolution". Biological Reviews. 92 (1): 446–473. doi:10.1111/brv.12239. PMID 26588818.
  14. The Air You're Breathing? A Diatom Made That
  15. Clark M A, Douglas M and Choi J (2018) Biology 2e, 23.4 "Ecology of Protists", OpenStax, Houston, Texas. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  16. Vallet, Marine; Baumeister, Tim U. H.; Kaftan, Filip; Grabe, Veit; Buaya, Anthony; Thines, Marco; Svatoš, Aleš; Pohnert, Georg (2019). "The oomycete Lagenisma coscinodisci hijacks host alkaloid synthesis during infection of a marine diatom". Nature Communications. 10 (1): 4938. doi:10.1038/s41467-019-12908-w. PMC 6821873. PMID 31666506.
  17. Beware the mixotrophs - they can destroy entire ecosystems 'in a matter of hours'
  18. Microscopic body snatchers infest our oceans - Phys.org
  19. Eiler A (December 2006). "Evidence for the Ubiquity of Mixotrophic Bacteria in the Upper Ocean: Implications and Consequences". Appl Environ Microbiol. 72 (12): 7431–7. doi:10.1128/AEM.01559-06. PMC 1694265. PMID 17028233.
  20. Katechakis A, Stibor H (July 2006). "The mixotroph Ochromonas tuberculata may invade and suppress specialist phago- and phototroph plankton communities depending on nutrient conditions". Oecologia. 148 (4): 692–701. Bibcode:2006Oecol.148..692K. doi:10.1007/s00442-006-0413-4. PMID 16568278. S2CID 22837754.
  21. Schoemann, Véronique; Becquevort, Sylvie; Stefels, Jacqueline; Rousseau, Véronique; Lancelot, Christiane (1 January 2005). "Phaeocystis blooms in the global ocean and their controlling mechanisms: a review". Journal of Sea Research. Iron Resources and Oceanic Nutrients - Advancement of Global Environmental Simulations. 53 (1–2): 43–66. Bibcode:2005JSR....53...43S. CiteSeerX 10.1.1.319.9563. doi:10.1016/j.seares.2004.01.008.
  22. "Welcome to the Phaeocystis antarctica genome sequencing project homepage".
  23. DiTullio, G. R.; Grebmeier, J. M.; Arrigo, K. R.; Lizotte, M. P.; Robinson, D. H.; Leventer, A.; Barry, J. P.; VanWoert, M. L.; Dunbar, R. B. (2000). "Rapid and early export of Phaeocystis antarctica blooms in the Ross Sea, Antarctica". Nature. 404 (6778): 595–598. doi:10.1038/35007061. PMID 10766240. S2CID 4409009.
  24. J, Stefels; L, Dijkhuizen; WWC, Gieskes (20 July 1995). "DMSP-lyase activity in a spring phytoplankton bloom off the Dutch coast, related to Phaeocystis sp. abundance" (PDF). Marine Ecology Progress Series. 123: 235–243. Bibcode:1995MEPS..123..235S. doi:10.3354/meps123235.
  25. Decelle, Johan; Simó, Rafel; Galí, Martí; Vargas, Colomban de; Colin, Sébastien; Desdevises, Yves; Bittner, Lucie; Probert, Ian; Not, Fabrice (30 October 2012). "An original mode of symbiosis in open ocean plankton". Proceedings of the National Academy of Sciences. 109 (44): 18000–18005. Bibcode:2012PNAS..10918000D. doi:10.1073/pnas.1212303109. ISSN 0027-8424. PMC 3497740. PMID 23071304.
  26. Mars Brisbin, Margaret; Grossmann, Mary M.; Mesrop, Lisa Y.; Mitarai, Satoshi (2018). "Intra-host Symbiont Diversity and Extended Symbiont Maintenance in Photosymbiotic Acantharea (Clade F)". Frontiers in Microbiology. 9: 1998. doi:10.3389/fmicb.2018.01998. ISSN 1664-302X. PMC 6120437. PMID 30210473.
  27. Stoecker, D.K.; Hansen, P.J.; Caron, D.A.; Mitra, A. (2017). "Mixotrophy in the marine plankton". Annual Review of Marine Science. 9: 311–335. Bibcode:2017ARMS....9..311S. doi:10.1146/annurev-marine-010816-060617. PMID 27483121.
  28. Mitra, A; Flynn, KJ; Tillmann, U; Raven, J; Caron, D; et al. (2016). "Defining planktonic protist functional groups on mechanisms for energy and nutrient acquisition; incorporation of diverse mixotrophic strategies". Protist. 167 (2): 106–20. doi:10.1016/j.protis.2016.01.003. PMID 26927496.
  29. Dawson, Scott C; Paredez, Alexander R (2013). "Alternative cytoskeletal landscapes: cytoskeletal novelty and evolution in basal excavate protists". Current Opinion in Cell Biology. 25 (1): 134–141. doi:10.1016/j.ceb.2012.11.005. PMC 4927265. PMID 23312067.
  30. Atkinson, A.; Polimene, L.; Fileman, E.S.; Widdicombe, C.E.; McEvoy, A.J.; Smyth, T.J.; Djeghri, N.; Sailley, S.F.; Cornwell, L.E. (2018). ""Comment. What drives plankton seasonality in a stratifying shelf sea? Some competing and complementary theories"]" (PDF). Limnology and Oceanography. 63 (6): 2877–2884. Bibcode:2018LimOc..63.2877A. doi:10.1002/lno.11036.
  31. Singleton, Paul (2006). Dictionary of Microbiology and Molecular Biology, 3rd Edition, revised. Chichester, UK: John Wiley & Sons. pp. 32. ISBN 978-0-470-03545-0.
  32. David J. Patterson. "Amoebae: Protists Which Move and Feed Using Pseudopodia". Tree of Life web project.
  33. "The Amoebae". The University of Edinburgh. Archived from the original on 10 June 2009.
  34. A Dictionary of Biology, 2004, accessed 2011-01-01.
  35. Patterson, David J. (2000) "Flagellates: Heterotrophic Protists With Flagella" Tree of Life.
  36. Lauga, Eric; Thomas R Powers (25 August 2009). "The hydrodynamics of swimming microorganisms". Reports on Progress in Physics. 72 (9): 096601. arXiv:0812.2887. Bibcode:2009RPPh...72i6601L. doi:10.1088/0034-4885/72/9/096601. S2CID 3932471.
  37. Guiry MD (October 2012). "How many species of algae are there?". Journal of Phycology. 48 (5): 1057–63. doi:10.1111/j.1529-8817.2012.01222.x. PMID 27011267. S2CID 30911529.
  38. Guiry, M.D.; Guiry, G.M. (2016). "Algaebase". www.algaebase.org. Retrieved 20 November 2016.
  39. D. Thomas (2002). Seaweeds. Life Series. Natural History Museum, London. ISBN 978-0-565-09175-0.
  40. Hoek, Christiaan; den Hoeck, Hoeck Van; Mann, David; Jahns, H.M. (1995). Algae : an introduction to phycology. Cambridge University Press. p. 166. ISBN 9780521316873. OCLC 443576944.
  41. Starckx, Senne (31 October 2012) A place in the sun - Algae is the crop of the future, according to researchers in Geel Flanders Today, Retrieved 8 December 2012
  42. Duval, B.; Margulis, L. (1995). "The microbial community of Ophrydium versatile colonies: endosymbionts, residents, and tenants". Symbiosis. 18: 181–210. PMID 11539474.
  43. Mandoli, DF (1998). "Elaboration of Body Plan and Phase Change during Development of Acetabularia: How Is the Complex Architecture of a Giant Unicell Built?". Annual Review of Plant Physiology and Plant Molecular Biology. 49: 173–198. doi:10.1146/annurev.arplant.49.1.173. PMID 15012232. S2CID 6241264.
  44. Pierre Madl; Maricela Yip (2004). "Literature Review of Caulerpa taxifolia". BUFUS-Info. 19 (31).
  45. Treguer, P.; Nelson, D. M.; Van Bennekom, A. J.; Demaster, D. J.; Leynaert, A.; Queguiner, B. (1995). "The Silica Balance in the World Ocean: A Reestimate". Science. 268 (5209): 375–9. Bibcode:1995Sci...268..375T. doi:10.1126/science.268.5209.375. PMID 17746543. S2CID 5672525.
  46. Nelson, David M.; Tréguer, Paul; Brzezinski, Mark A.; Leynaert, Aude; Quéguiner, Bernard (1995). "Production and dissolution of biogenic silica in the ocean: Revised global estimates, comparison with regional data and relationship to biogenic sedimentation". Global Biogeochemical Cycles. 9 (3): 359–372. Bibcode:1995GBioC...9..359N. doi:10.1029/95GB01070.
  47. Malviya, Shruti; Scalco, Eleonora; Audic, Stéphane; Vincent, Flora; Veluchamy, Alaguraj; Poulain, Julie; Wincker, Patrick; Iudicone, Daniele; De Vargas, Colomban; Bittner, Lucie; Zingone, Adriana; Bowler, Chris (2016). "Insights into global diatom distribution and diversity in the world's ocean". Proceedings of the National Academy of Sciences. 113 (11): E1516–E1525. Bibcode:2016PNAS..113E1516M. doi:10.1073/pnas.1509523113. PMC 4801293. PMID 26929361. S2CID 22035749.
  48. Tréguer, Paul; Bowler, Chris; Moriceau, Brivaela; Dutkiewicz, Stephanie; Gehlen, Marion; Aumont, Olivier; Bittner, Lucie; Dugdale, Richard; Finkel, Zoe; Iudicone, Daniele; Jahn, Oliver; Guidi, Lionel; Lasbleiz, Marine; Leblanc, Karine; Levy, Marina; Pondaven, Philippe (2018). "Influence of diatom diversity on the ocean biological carbon pump". Nature Geoscience. 11 (1): 27–37. Bibcode:2018NatGe..11...27T. doi:10.1038/s41561-017-0028-x. S2CID 134885922.
  49. Mahadevan, Amala; d'Asaro, Eric; Lee, Craig; Perry, Mary Jane (2012). "Eddy-Driven Stratification Initiates North Atlantic Spring Phytoplankton Blooms". Science. 337 (6090): 54–58. Bibcode:2012Sci...337...54M. doi:10.1126/science.1218740. PMID 22767922. S2CID 42312402.
  50. Cavicchioli, Ricardo; Ripple, William J.; Timmis, Kenneth N.; Azam, Farooq; Bakken, Lars R.; Baylis, Matthew; Behrenfeld, Michael J.; Boetius, Antje; Boyd, Philip W.; Classen, Aimée T.; Crowther, Thomas W.; Danovaro, Roberto; Foreman, Christine M.; Huisman, Jef; Hutchins, David A.; Jansson, Janet K.; Karl, David M.; Koskella, Britt; Mark Welch, David B.; Martiny, Jennifer B. H.; Moran, Mary Ann; Orphan, Victoria J.; Reay, David S.; Remais, Justin V.; Rich, Virginia I.; Singh, Brajesh K.; Stein, Lisa Y.; Stewart, Frank J.; Sullivan, Matthew B.; et al. (2019). "Scientists' warning to humanity: Microorganisms and climate change". Nature Reviews Microbiology. 17 (9): 569–586. doi:10.1038/s41579-019-0222-5. PMC 7136171. PMID 31213707. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  51. Wassilieff, Maggy (2006) "Plankton - Plant plankton", Te Ara - the Encyclopedia of New Zealand. Accessed: 2 November 2019.
  52. "King's College London - Lake Megachad". www.kcl.ac.uk. Retrieved 5 May 2018.
  53. Boyd, Philip W.; Claustre, Hervé; Levy, Marina; Siegel, David A.; Weber, Thomas (2019). "Multi-faceted particle pumps drive carbon sequestration in the ocean". Nature. 568 (7752): 327–335. Bibcode:2019Natur.568..327B. doi:10.1038/s41586-019-1098-2. PMID 30996317. S2CID 119513489.
  54. Zhang, D.; Wang, Y.; Cai, J.; Pan, J.; Jiang, X.; Jiang, Y. (2012). "Bio-manufacturing technology based on diatom micro- and nanostructure". Chinese Science Bulletin. 57 (30): 3836–3849. Bibcode:2012ChSBu..57.3836Z. doi:10.1007/s11434-012-5410-x.
  55. Behrenfeld, Michael J.; Doney, Scott C.; Lima, Ivan; Boss, Emmanuel S.; Siegel, David A. (2013). "Annual cycles of ecological disturbance and recovery underlying the subarctic Atlantic spring plankton bloom". Global Biogeochemical Cycles. 27 (2): 526–540. Bibcode:2013GBioC..27..526B. doi:10.1002/gbc.20050.
  56. Rousseaux, Cecile S.; Gregg, Watson W. (2015). "Recent decadal trends in global phytoplankton composition". Global Biogeochemical Cycles. 29 (10): 1674–1688. Bibcode:2015GBioC..29.1674R. doi:10.1002/2015GB005139.
  57. Arsenieff, L.; Simon, N.; Rigaut-Jalabert, F.; Le Gall, F.; Chaffron, S.; Corre, E.; Com, E.; Bigeard, E.; Baudoux, A.C. (2018). "First Viruses Infecting the Marine Diatom Guinardia delicatula". Frontiers in Microbiology. 9: 3235. doi:10.3389/fmicb.2018.03235. PMC 6334475. PMID 30687251.
  58. Kilias, Estelle S.; Junges, Leandro; Šupraha, Luka; Leonard, Guy; Metfies, Katja; Richards, Thomas A. (2020). "Chytrid fungi distribution and co-occurrence with diatoms correlate with sea ice melt in the Arctic Ocean". Communications Biology. 3 (1): 183. doi:10.1038/s42003-020-0891-7. PMC 7174370. PMID 32317738. S2CID 216033140. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  59. Rost, B. and Riebesell, U. (2004) "Coccolithophores and the biological pump: responses to environmental changes". In: Coccolithophores: From Molecular Processes to Global Impact, pages 99–125, Springer. ISBN 9783662062784.
  60. Wassilieff, Maggy (2006) "A coccolithophore", Te Ara - the Encyclopedia of New Zealand. Accessed: 2 November 2019.
  61. Hagino, K., Onuma, R., Kawachi, M. and Horiguchi, T. (2013) "Discovery of an endosymbiotic nitrogen-fixing cyanobacterium UCYN-A in Braarudosphaera bigelowii (Prymnesiophyceae)". PLoS One, 8(12): e81749. doi:10.1371/journal.pone.0081749.
  62. Gómez F (2012). "A checklist and classification of living dinoflagellates (Dinoflagellata, Alveolata)" (PDF). CICIMAR Océanides. 27 (1): 65–140. Archived from the original (PDF) on 27 November 2013.
  63. Stoecker DK (1999). "Mixotrophy among Dinoflagellates". The Journal of Eukaryotic Microbiology. 46 (4): 397–401. doi:10.1111/j.1550-7408.1999.tb04619.x. S2CID 83885629.
  64. Suggested Explanation for Glowing Seas--Including Currently Glowing California Seas National Science Foundation, 18 October 2011.
  65. Boltovskoy, Demetrio; Anderson, O. Roger; Correa, Nancy M. (2017). Handbook of the Protists. Springer, Cham. pp. 731–763. doi:10.1007/978-3-319-28149-0_19. ISBN 9783319281476.
  66. Anderson, O. R. (1983). Radiolaria. Springer Science & Business Media.
  67. Gast, R. J.; Caron, D. A. (1 November 1996). "Molecular phylogeny of symbiotic dinoflagellates from planktonic foraminifera and radiolaria". Molecular Biology and Evolution. 13 (9): 1192–1197. doi:10.1093/oxfordjournals.molbev.a025684. ISSN 0737-4038. PMID 8896371.
  68. Castro P, Huber ME (2010). Marine Biology (8th ed.). McGraw Hill. pp. 95. ISBN 978-0071113021.
  69. Hastings JW (1996). "Chemistries and colors of bioluminescent reactions: a review". Gene. 173 (1 Spec No): 5–11. doi:10.1016/0378-1119(95)00676-1. PMID 8707056.
  70. Haddock SH, Moline MA, Case JF (2009). "Bioluminescence in the sea". Annual Review of Marine Science. 2: 443–93. Bibcode:2010ARMS....2..443H. doi:10.1146/annurev-marine-120308-081028. PMID 21141672. S2CID 3872860.
  71. "Protozoa Infecting Gills and Skin". The Merck Veterinary Manual. Archived from the original on 3 March 2016. Retrieved 4 November 2019.
  72. Brand, Larry E.; Campbell, Lisa; Bresnan, Eileen (2012). "Karenia: The biology and ecology of a toxic genus". Harmful Algae. 14: 156–178. doi:10.1016/j.hal.2011.10.020.
  73. Buskey, E.J. (1995). "Growth and bioluminescence of Noctiluca scintillans on varying algal diets". Journal of Plankton Research. 17 (1): 29–40. doi:10.1093/plankt/17.1.29.
  74. Panno, Joseph (14 May 2014). The Cell: Evolution of the First Organism. Infobase Publishing. ISBN 9780816067367.
  75. Bertrand, Jean-Claude; Caumette, Pierre; Lebaron, Philippe; Matheron, Robert; Normand, Philippe; Sime-Ngando, Télesphore (26 January 2015). Environmental Microbiology: Fundamentals and Applications: Microbial Ecology. Springer. ISBN 9789401791182.
  76. Madigan, Michael T. (2012). Brock Biology of Microorganisms. Benjamin Cummings. ISBN 9780321649638.
  77. Yaeger, Robert G. (1996). Protozoa: Structure, Classification, Growth, and Development. NCBI. ISBN 9780963117212. PMID 21413323. Retrieved 23 March 2018.
  78. Wassilieff, Maggy (2006) "Plankton - Animal plankton", Te Ara - the Encyclopedia of New Zealand. Accessed: 2 November 2019.
  79. Varea, C.; Aragon, J.L.; Barrio, R.A. (1999). "Turing patterns on a sphere". Physical Review E. 60 (4): 4588–92. Bibcode:1999PhRvE..60.4588V. doi:10.1103/PhysRevE.60.4588. PMID 11970318.
  80. Hemleben, C.; Anderson, O.R.; Spindler, M. (1989). Modern Planktonic Foraminifera. Springer-Verlag. ISBN 978-3-540-96815-3.
  81. Foraminifera: History of Study, University College London. Retrieved: 18 November 2019.
  82. Advances in Microbial Ecology, Volum 11
  83. Bernhard, J. M.; Bowser, S.M. (1999). "Benthic Foraminifera of dysoxic sediments: chloroplast sequestration and functional morphology". Earth-Science Reviews. 46 (1): 149–165. Bibcode:1999ESRv...46..149B. doi:10.1016/S0012-8252(99)00017-3.
  84. Matz, Mikhail V.; Tamara M. Frank; N. Justin Marshall; Edith A. Widder; Sonke Johnsen (9 December 2008). "Giant Deep-Sea Protist Produces Bilaterian-like Traces" (PDF). Current Biology. Elsevier Ltd. 18 (23): 1849–1854. doi:10.1016/j.cub.2008.10.028. PMID 19026540. S2CID 8819675.
  85. Gooday, A. J.; Aranda da Silva, A.; Pawlowski, J. (1 December 2011). "Xenophyophores (Rhizaria, Foraminifera) from the Nazaré Canyon (Portuguese margin, NE Atlantic)". Deep-Sea Research Part II: Topical Studies in Oceanography. The Geology, Geochemistry, and Biology of Submarine Canyons West of Portugal. 58 (23–24): 2401–2419. Bibcode:2011DSRII..58.2401G. doi:10.1016/j.dsr2.2011.04.005.
  86. Biodegradable glitter and pollution-eating microalgae: the new materials inspired by nature Horizon, 28 May 2020.
  87. Aguirre, L.E., Ouyang, L., Elfwing, A., Hedblom, M., Wulff, A. and Inganäs, O. (2018) "Diatom frustules protect DNA from ultraviolet light". Scientific reports, 8(1): 1–6. doi:10.1038/s41598-018-21810-2. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  88. De Tommasi, E., Congestri, R., Dardano, P., De Luca, A.C., Managò, S., Rea, I. and De Stefano, M. (2018) "UV-shielding and wavelength conversion by centric diatom nanopatterned frustules". Nature: Scientific Reports, 8(1): 1–14. doi:10.1038/s41598-018-34651-w. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  89. Kachovich, S., Sheng, J. and Aitchison, J.C., 2019. Adding a new dimension to investigations of early radiolarian evolution. Scientific reports, 9(1), pp.1-10. doi:10.1038/s41598-019-42771-0. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  90. Monteiro, F.M., Bach, L.T., Brownlee, C., Bown, P., Rickaby, R.E., Poulton, A.J., Tyrrell, T., Beaufort, L., Dutkiewicz, S., Gibbs, S. and Gutowska, M.A. (2016) "Why marine phytoplankton calcify". Science Advances, 2(7): e1501822. doi:10.1126/sciadv.1501822. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  91. Booth, B.C. and Marchant, H.J. (1987) "Parmales, a new order of marine chrysophytes, with descriptions of three new genera and seven new species". Journal of Phycology, 23: 245–260. doi:10.1111/j.1529-8817.1987.tb04132.x.
  92. Kuwata, A., Yamada, K., Ichinomiya, M., Yoshikawa, S., Tragin, M., Vaulot, D. and Lopes dos Santos, A. (2018) "Bolidophyceae, a sister picoplanktonic group of diatoms – a review". Frontiers in Marine Science, 5: 370. doi:10.3389/fmars.2018.00370. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  93. Gafar, N.A., Eyre, B.D. and Schulz, K.G. (2019) "A comparison of species specific sensitivities to changing light and carbonate chemistry in calcifying marine phytoplankton". Nature: Scientific Reports, 9(1): 1–12. doi:10.1038/s41598-019-38661-0. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.

Further references

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.