Invasion History

First Non-native North American Tidal Record: 1944
First Non-native West Coast Tidal Record: 1944
First Non-native East/Gulf Coast Tidal Record:

General Invasion History:

Sargassum muticum was first described from Wakayama Prefecture, Japan, by Yendo in 1907. Its native range extends from the East China Sea (China and South Korea) to the coasts of Japan and the north shore of Hokkaido, and the Sakhalin and Kurile Islands, Russia (Critichley 1983b; Eneglen et al. 2015). It is introduced in the Eastern Pacific and Eastern Atlantic and genetic studies indicate that these populations probably originated from the Seto Inland Sea, Japan (Cheang et al. 2010). In North America, S. muticum was first introduced to Vancouver Island, British Columbia, where it was discovered in 1944, probably introduced with Pacific Oysters (Crassostrea gigas) from Japan. By 1977, it had spread north to Ketchikan, Alaska (Scagel et al. 1956; Scagel et al. 1989). The pattern of spread/discovery to the south was rather jumpy, with first records in Coos Bay, Oregon in 1947; Mission Bay, California (CA) in 1959; San Francisco Bay, CA in 1963; and Los Angeles, CA in 1973. By the 2000s it had colonized most of the Pacific coast of Baja California (Scagel et al. 1956; Scagel et al. 1989; Engelen et al. 2015). In Europe, it was first collected in Southeast England in 1971 and rapidly spread north and south, reaching Norway by 1988, Spain by 1985, and Morocco by 2012 (Farnham 1980; Runess 1989; Fernandez et al. 1990; Sabour et al. 2013). In 1980, it was first collected in lagoons on the Mediterranean coast of France and by 1992 it was established in the Lagoon of Venice (Knoepfller-Peguy et al. 1985; Curiel et al. 1998), but its distribution in the Mediterranean Sea remains localized (Engelen et al. 2015; Thibaut et al. 2015). The large size of this seaweed, its rapid spread, and high abundances in many locations, have led to the extensive study of its ecology and impacts (Engelen et al. 2015).

North American Invasion History:

Invasion History on the West Coast:

Sargassum muticum was first collected on the West Coast of North America, at White Rock and Buccaneer Bay, north of Vancouver, British Columbia in 1944, in sites where Pacific Oysters where cultivated. It was initially mistaken for the native seaweed Cytoseira geminata. By 1952, it was found at many sites in the Strait of Georgia and Puget Sound, including the San Juan Islands and the Strait of Juan de Fuca (Scagel 1956). The pattern of discovery along the coast of Oregon and Washington suggests several independent introductions with oyster transplants, with subsequent spread by drifting weed and fouled boats. It was found in Willapa Bay, Washington in 1953; Oceanside Beach (near Tillamook Bay), Oregon, in 1951; and Coos Bay, OR in 1947 (Scagel 1956). Spread to the north was slower, but it was found near Ketchikan, Alaska in 1974-1977 (Scagel et al. 1989; Engelen et al. 2015; University of Alaska Southeast Herbarium 2016), and Haida Gwaii, British Columbia in 1981 (Sloan and Bartier 2004).

The first record of S. muticum in California is a report of its occurrence in Mission Bay, San Diego in 1958, which looks like a sudden jump (Stewart 1991), possibly by ship transport. Other records are consistent with a progressive march to the south: Crescent City Harbor in 1963 (Norton 1981, Cohen 2005); Humboldt Bay in 1965 (Dawson 1965, cited by Boyd et al. 2002); Tomales Bay and San Francisco Bay in 1973 (Cohen and Carlton 1995); Monterey Bay in 1977 (Carlton et al. 1996; Cohen 2005); Santa Barbara in 1977 (Cohen 2005); Los Angeles in 1973; (Norton 1981); and San Diego Bay in 1969 (Cohen 2005). The rapid spread is surprising, since S. muticum requires sheltered waters for establishment (Norton 1981). By 1973, it was already established in Ensenada, Mexico (Norton 1981) and by 1984 it had reached Punta Abreojos, Baja California, Mexico (Espinosa 1990). Overall, in 70 years since its invasion on the West coast, it has colonized ~30 degrees of latitude and a span of 4000 km of coastline (Engelen et al. 2015).

Invasion History Elsewhere in the World:

In the Northeast Atlantic, S. muticum was first found on Portsea Island, Portsmouth, England in 1971, and rapidly spread through the Solent, between the south coast of England and the Isle of Wight (Farnham 1980). It colonized the French side of the English Channel by 1977 (Gruet 1977, cited by Farnham 1980). This seaweed reached the Netherlands by 1980, the German portion of the Wadden Sea by 1983 (Buschbaum et al. 2012), and the western mouth of the Limfjord, Denmark, by 1984 (Staer et al. 2000). Sargassum muticum colonized the Kattegat in Denmark and Sweden, the westernmost portion of the Baltic Sea, in 1993-1996, probably reaching the lower limit of its salinity tolerance (Karlsson and Loo 1999; Thomsen et al. 2007). Moving north, S. muticum spread into Norwegian and Swedish waters of the Skaggerak in 1987-1988 (Rueness 1989; Karsson and Loo 1999) and north up the west coast of Norway, to Bergen in 1989 (Hopkins 202) and Sogn og Fjordane (Bjærke and Fredriksen 2005). To the south, this seaweed reached Roscoff, at the tip of Brittany in 1980 (Belsher and Pommellec 1988), and reached the Basque Province of Spain by 1985 (Fernandez et al. 1990), and Portugal by 1989 (Chainho et al. 2015). In 2012, S. muticum was collected at several sites on the Atlantic coast of Morocco. The southernmost was Oualidia at 32°N (Sabour et al. 2013). In ~40 years, S. muticum has spread over 39 degrees of latitude and a 5600 km span of coastline (Engelen et al. 2015).

In contrast with its spread on the West Coasts of Europe and North America, Sargassum muticum's spread in the Mediterranean has been limited, spotty, and punctuated by extinctions. In 1985-1990, S. muticum was found in 14 lagoons on the French Mediterranean Coast, but in surveys in 2012 established populations were found in only four lagoons (Thibaut et al. 2015). In 1992, established populations of the seaweed were found in the Venice Lagoon. Dense populations occur on wharves and breakwaters (Curiel et al. 1998). Sargassum muticum was found in a lagoon on the east coast of Corsica, but was not seen in later surveys (Thibaut et al. 2015). Some floating specimens of this seaweed have been found in the open Mediterranean off Spain, but S. muticum has only become established in lagoons with limited exchange (Engelen et al. 2015).

Culture of the Pacific Oyster (Crassostrea gigas) has been the initial vector for the three major invasions of S. muticum in the Northeast Pacific and Atlantic, and the Mediterranean (Scagel 1956; Farnham 1980; Knoepffler-Peguy et al. 1985). Regional and local dispersal could occur by hull fouling of commercial ships or recreational boats and by the natural dispersal of floating mats of seaweed (Engelen et al. 2015).


Description

Sargassum muticum usually grows upright in the water column, attached by a solid, conical, spongy holdfast about 50 mm across. The holdfast gives rise to a tapered axis up to 50 mm in height. In young plants, leaves, up to 200 mm long, arise from the main axis. These basal leaves have a well-defined mid-rib. In older plants, the main axis may branch, making the plant bushier. Lateral shoots grow off the main axis in spirals with scale-like triangular leaves at the base, which protect the growing bud. The main axis is perennial, while the lateral shoots die off after the growing season. Primary lateral stems are cord-like, usually growing to 1-3 m, but sometimes 6-10 m. In populations on the West Coast of Ireland, mature plants had 40-100 primary lateral branches (Baer and Stengel 2010), while a population in the Netherlands had 1-16 (Critchley et al. 1987). Secondary lateral stems arise off the primary laterals, branching off in the axil of a leaf. As the lateral stems grow, they become twisted due to water movement. Leaves from the primary and secondary laterals, produced in the winter, are small, less than 200 mm long, thick, and have a weak mid-rib. During the summer, the fronds become fertile and the new leaves are thinner, narrow, and lacking a midrib. They have an uneven outline, resembling holly leaves. The lateral branches bear air-bladders and reproductive receptacles in the axils of the leaves. The air-bladders are spherical to pear-shaped, on stalks, about 3 mm in diameter. The receptacles vary in shape from elliptical to cylindrical to spindle-shaped and are usually 20--30 mm long (sometimes up to 60 mm). The plants are yellow-green to olive-brown. This description is based on: Abbott and Hollenberg 1976; Critchley 1983a; Cohen 2005; and Engelen et al. 2015.

Sargassum muticum is one of several similar species known in Japan, and its taxonomic history is complex. It was initially described as a variety of S. kjellmanianum, which is now regarded as a synonym of S. miyabe (Critchley 1983b; Cheang et al. 2010; Engelen et al. 2015). The genetic diversity of introduced populations in North America and Europe is low and all populations are close to S. muticum populations from western and central Japan (Cheang et al. 2010).


Taxonomy

Taxonomic Tree

Kingdom:   Plantae
Phylum:   Phaeophycophyta
Class:   Phaeophyceae
Order:   Fucales
Family:   Sargassaceae
Genus:   Sargassum
Species:   muticum

Synonyms

Sargassum kjellmanianum f. muticum (Yendo, 1907)
Sargassum muticum (None, None)

Potentially Misidentified Species

Cytoseira geminata
Native Northeast Pacific form, morphological details given by Scagel (1956)

Sargassum horneri
Northwest Pacific species, introduced to southern California and Mexico

Sargassum miyabei
Northwest Pacific species, morphological details given by Critchley (1983)

Ecology

General:

Sargassum muticum is monoecious and mature plants produce receptacles, which contain both male and female conceptacles for the production of sperm and eggs (Bold and Wynne 1978; Engelen et al. 2015). Receptacles can constitute 24-55% of the plant's biomass, but this reproductive output is fairly typical for fucoid seaweeds (Engelen et al. 2015). Gamete release is synchronized in a semilunar, 14-day cycle, with release peaking at the full and new moons, just after spring tides (Engelen et al. 2008; Engelen et al. 2015), though with some variation with local cues (Monteiro et al. 2009b). Eggs are fertilized in the receptacles, but are retained, attached to the receptacle surface for a few days and are then released as germlings that sink quickly. The propagules usually settle within a few meters of the parents, but can be found up to 1.3 km away. They could potentially disperse over longer distances, since they retain the ability to settle for 49 days (Deysher and Norton 1982). Each receptacle releases ~ 300 propagules and a small plant can release up to 500,000 propagules (Norton and Deysher 1989, cited by Engelen et al. 2015). The fertilized zygotes average about 0.25 mm in diameter (Deysher and Norton 1982). Germlings have comparatively large, basal leaves, with developing lateral axes (Critchley 1983a). Breeding seasons vary geographically, peaking in spring-summer in Northern Europe and Northwestern North America, but with a longer season at southern locations such as Portugal, southern California, and Mexico (Deysher 1984; Aguilar-Rosas and Galindo 1990; Engelen et al. 2015). After releasing their propagules, adult plants lose their fronds and disintegrate, except for the perennial holdfast, which gives rise to new fronds in the next growing season (Critchley 1983a; Engelen et al. 2015). This dormant period is usually in autumn in the northern part of the range, but occurs in summer-fall in southern Portugal, California, and the Venice Lagoon (Deysher 1984; Sfriso and Facca 2013; Engelen et al. 2015).

Sargassum muticum grows over a wide latitudinal range, from cold-temperate to subtropical conditions, from temperatures below 0°C in Sweden (Karlsson and Loo 1999) to 30°C in the Venice Lagoon (Sfriso and Facca 2013). In the laboratory, germlings survived at salinities as low as 6.8 PSU, but with minimal growth (Hales and Fletcher 1989; Karlson and Loo 1999; Steen 2004; Thomsen et al. 2007). Germlings are more sensitive to temperature and salinity than adult plants and grow successfully between 15 and 25°C and 15-35 PSU (Hales and Fletcher 1989; Steen 2004). Early germlings (2 weeks old) showed increasing growth at 9 to 44 µE m-3s-1 and had decreased growth at 88 µE m-3s-1, while older germlings grew well at 18 to 88 µE m-3s-1, and were more tolerant of high light (Hales and Fletcher 1989). Sargassum muticum is not a good competitor at low light and has a strategy of growing fast to form a canopy in shallow water. It can adjust its pattern of growth and branching according to the density of its neighbors, permitting it to form very dense populations (Engelen et al. 2015).

In its native range in Asia, S. muticum occurs on rocky shores from the lower intertidal to 4 m depth and is not known as an aggressive colonizer of artificial structures (Engelen et al. 2015). It is generally associated with sheltered locations. Plants transplanted to exposed locations suffered increased breakage and lower growth rates (Viejo et al. 1995). In invaded habitats, it occurs on pilings, floats, marinas, in canals, aquaculture cages, marinas, breakwaters, oyster beds, and eelgrass beds (Belsher and Pommellec 1988; den Hartog 1997; Harries 2007; Kraan 2008; Engelen et al. 2015). In Strangford Lough, Northern Ireland, S. muticum has colonized soft-sediment habitats by 'stone-walking' on rock fragments and shells and spreading by water movements (Strong et al. 2006). Adult plants, dislodged by waves or disturbance, or fragments of degenerating plants with fertile receptacles, form floating mats (Norton 1981). Sargassum muticum, like other fucoid seaweeds, produces organic compounds for chemical defense against microbes, epiphytes, and grazers. Phenolic compounds are greatest during the reproductive period, possibly for protection of the receptacles (Plouguerne et al. 2006; Plougerne et al. 2008). In spite of these compounds, S. muticum supports a dense community of epiphytes and epifauna, and is grazed by a variety of invertebrates, with varying degrees of preference (Norton and Benson 1983; Monteiro et al. 2009; Strong et al. 2009; Baer & Stengel 2010; Cacabelos et al. 2010; Gestoso et al. 2010).

Consumers:

snails, amphipods, sea urchins

Competitors:

Native seaweeds, seagrasses

Trophic Status:

Primary Producer

PrimProd

Habitats

General HabitatMarinas & DocksNone
General HabitatRockyNone
General HabitatOyster ReefNone
General HabitatGrass BedNone
General HabitatCanalsNone
General HabitatUnstructured BottomNone
Salinity RangeMesohaline5-18 PSU
Salinity RangePolyhaline18-30 PSU
Salinity RangeEuhaline30-40 PSU
Tidal RangeSubtidalNone
Tidal RangeLow IntertidalNone
Vertical HabitatEpibenthicNone
Vertical HabitatLittoralNone

Life History


Tolerances and Life History Parameters

Minimum Temperature (ºC)0Field observations, Sweden (Karlsson and Loo 1999)
Maximum Temperature (ºC)30Highest tested, lab experiments (Hales and Fletcher 1989)
Minimum Salinity (‰)10Lab experiments- Growth greatly reduced compared to 27-34 PSU (Hales and Fletcher 1989)
Maximum Salinity (‰)35Highest tested, lab experiments (Hales and Fletcher 1989)
Minimum Reproductive Temperature15Experimental (Kerrison and Le 2016)
Maximum Reproductive Temperature25Experimental (Kerrison and Le 2016)
Minimum Reproductive Salinity20Minimum for fertilization. Germlings can tolerate brief exposure to 5 PSU, especially in 2nd and 3rd weeks after fertilization (Steen 1984)
Maximum Reproductive Salinity40Experimental, germlings in receptacle (Kerrison and Le2016)
Minimum Length (mm)300Moyrus, Ireland, stunted plants at exposed site, but 100% fertile (Baer and Stengel 2010)
Maximum Length (mm)10,000Engelen et al. 2015
Broad Temperature RangeNoneCold temperate-Warm temperate
Broad Salinity RangeNoneMesohaline-Euhaline

General Impacts

Sargassum muticum is one of many Sargassum species in its native Asian range, and is not especially prominent or aggressive there. In the Northeast Pacific and Northeast Atlantic, however, it has spread rapidly along coastlines and has become a major component of algal communities (Engelen et al. 2015). Economic impacts to beach use, boating, power plants, fisheries, and aquaculture, were reported in the early years of the invasion in England and France. These impacts led to a number of unsuccessful eradication attempts (Critchley et al. 1986; Belsher and Pommellec 1989). Ecological impacts of S. muticum have been extensively studied, particularly along the coasts of northern Europe, northern Spain and Portugal, and the San Juan Islands, in northern Puget Sound, Washington. These include competition for space, light, and nutrients, alterations of habitat and associated epibiota, and effects on grazers (Schaffelke and Hewitt 2007; Engelen et al. 2015).

Economic Impacts

Early accounts of the spread of Sargassum muticum on the West Coast of North America (Scagel 1956; Norton 1981) make no mention of adverse economic impacts. A later survey, focused on Sargassum horneri and Undaria pinnatifida refers to S.muticum as 'naturalized' and does not refer to economic impacts (Kaplanis et al. 2016). In Southeastern England and Atlantic France, extensive impacts, including fouling of boats, fishing nets, aquaculture equipment, and power plant intakes, was reported, together with large floating mats in nearshore waters and rotting masses piling up on beaches (Critchley et al. 1986; Belsher and Pommellec 1988). In 1974-1975, an eradication was attempted in England with large volunteer groups, hand-picking intertidal plants. The volunteers collected 31 metric tons, but missed germlings and juvenile plants and eradication was found to be impractical. Trawls, cutting machines, and suction machines were tested as alternatives to hand-picking, but considered un-economical. Sargassum muticum was considered unsuitable for industrial and food products (Critchley et al. 1986), although it is eaten in Korea. A recent review of potential economic uses found that use of 'wild' plants in Europe for biofuel, food, or fertilizer was unsafe because of the accumulation of heavy metals and high ash content. However, S. muticum contains fucoxanthins, antioxidants, and other compounds of pharmaceutical interest (Milledge et al. 2106).

Ecological Impacts

Competition- Interactions between S. muticum and native biota, in many different marine communities have been studied, including rocky shores, tidepools, kelp beds, seagrass meadows, and soft-bottom communities (Schaffelke and Hewitt 2007; Engelen et al. 2008). Competitive interactions have been especially noted with the algae of the closely related perennial genera Cytoseira and Halidrys spp. Sargassum’s semi-perennial life-style, dying back to the holdfast, in unfavorable seasons may be a competitive advantage. In Portugal, its dying fronds denied space to C. humlis (Engelen and Santos 2009). In Denmark, S. muticum outgrew Halidrys siliqua, possibly because it did not need to invest in structural strength for winter survival (Wernberg et al. 2000). Sargassum muticum's bushy growth form and buoyant branches permit it to form a canopy and shut out light and occupy space during the most favorable time for growth (Curiel et al. 1998; Britton-Simmons 2004; Harries et al. 2007; Salvaterra et al. 2013). However, S. muticum often seems to require disturbance to invade established communities (de Wreede 1983, cited by Schaffelke and Hewitt 2007; Strong and Dring 2011; Bertocci et al. 2014). Sargassum muticum invaded rocky areas off California and in the San Juan Islands after die-offs or experimental removal of kelps (Ambrose and Nelson 1982; Britton-Simmons and Abbott 2008).

Habitat Change- When S. muticum replaces native seaweeds and seagrasses, it can alter the structure of native habitats. In many cases, S. muticum supports a similar or a more abundant and diverse epiphytic and epifaunal community than the native flora (Viejo 1999; Buschbaum et al. 2006; Engelen et al. 2015). Its ability to colonize bare, soft substrate by 'stone-walking' on shells and stones, and to colonize Eelgrass (Zostera marina) beds greatly increases the structural complexity of these habitats (den Hartog 1997; Strong et al. 2011; DeAmicis et al. 2015). It is used as habitat by fishes, crustaceans, and cuttlefish (Engelen et al. 2015; Stiger-Pouvreau and Thouzeau 2015).

Food/Prey- Herbivores vary considerably in their response to S. muticum. Strong avoidance was noted for sea urchins (Strongylocentrotus droebachiensis) in northern Puget Sound (Britton-Simmons 2004) and Psammechinus miliaris in Denmark (Pedersen et al. 2005). Other grazers showed lesser levels of avoidance, or no preference (Monteiro et al. 2009; Cacabelos et al. 2010; Engelen et al. 2011). Some grazers appear to feed heavily on S. muticum, such as the amphipod Dexamine spinosa in Strangford Lough, Northern Ireland (Strong et al. 2009), and the snail Lacuna vincta in Northern Puget Sound. The latter seems to have developed its preference in the last 30 years (Britton-Simmons et al. 2011). In many of the systems where it has been studied, S. muticum seems to be under light grazing pressure (Norton and Benson 1983; Pedersen et al. 2005; Monteiro et al. 2009; Cacabelos et al. 2010; Engelen et al. 2011; Mabey et al. 2022). However, because of its rapid growth, large biomass, and seasonal die-offs, it is a major contributor to primary production in many intertidal and subtidal systems, and to subtidal detritus and intertidal wrack (Rossi et al. 2009; Olabarria et al. 2010; Vaz-Pinto et al. 2014; Engelen et al. 2015).


Regional Impacts

MED-VIINoneEcological ImpactCompetition
In the Venice Lagoon, S. muticum was found to compete with the native red, brown, and green macroalgae by forming a canopy and shutting out light (Curiel et al. 1998).
NEA-IINoneEcological ImpactCompetition
Sargassum muticum has been reported to displace native algae in the Limfjord, Denmark, (Stæhr et al. 2000, cited by Schaffelke and Hewitt 2007). Its advantage over the native, perennial Halidrys siliqua may be due to its semi-perennial lifestyle, not having to invest in a sturdier structure for winter survival, and so is capable of more rapid growth in spring and summer (Wernberg et al. 2000). On the Isle of Cumbrae, Scotland, S. muticum was found to displace the native Dictyota dichotoma, probably through competition for light and substrate (Harries et al. 2007). In Strangford Lough, Northern Ireland, competition with the native Saccharina latissima was not seen. Instead, growth of S. muticum was slower in single-species plots, due to intraspecific competition. Disturbance and removal of the native seaweed was considered responsible for the expansion of S. muticum (Strong and Dring 2011). In the German Wadden Sea, competition with the native algae Polysiphonia nigrescens, Antithamnion plumula and Elachista fucicola, and also with settling Pacific Oysters (Crassostrea gigas) was seen. Reduced oyster settlement could affect epibenthic communities by reducing by limiting the expansion of hard substrate (Lang and Buschbaum 2010).
NEP-IIIAlaskan panhandle to N. of Puget SoundEcological ImpactCompetition
The invasion of Sargassum muticum resulted in displacement of native algae in the San Juan Islands, northern Puget Sound. A removal experiment resulted in recovery of native kelps (Britton-Simmons 2004). Modeling and experiments indicated that S. muticum invasions required a combination of disturbance and high propagule pressure (Britton-Simmons et al. 2008). In British Columbia, S. muticum rapidly colonized cleared areas, followed by decreased recruitment of native seaweeds (de Wreede 1983, cited by Schaffelke and Hewitt 2007). At low levels of abundance, S. muticum had few impacts, but at high levels, S. muticum excludes natives through competition for light (primarily) and space, resulting in reduced diversity and productivity. The impact increased with S. muticum density in a non-linear fashion (White and Shurin 2011).
NEP-IIIAlaskan panhandle to N. of Puget SoundEcological ImpactFood/Prey
There were fewer Green Sea Urchins (Strongylocentrotus droebachiensis) at invaded sites, apparently because they found S. muticum unpalatable (Britton-Simmons 2004). However, the snail Lacuna vincta was 2-9X more abundant on S. muticum than on native algae. This preference seems to have been acquired in the last 30 years (Britton-Simon et al. 2011).
P292_CDA_P292 (San Juan Islands)Ecological ImpactCompetition
The invasion of Sargassum muticum resulted in displacement of native algae, in the San Juan Islands, northern Puget Sound. A removal experiment resulted in recovery of native kelps (Britton-Simmons 2004). Modeling and experiments indicated that S. muticum invasions required a combination of disturbance and high propagule pressure (Britton-Simmons et al. 2008).
P292_CDA_P292 (San Juan Islands)Ecological ImpactFood/Prey
Sargassum muticum was grazed by a high abundance, but low diversity of grazers, compared to native seaweeds. Common grazers were the amphipods Peramphithoe mea, Aoroides columbiae, Caprella laeviscula and Ischyocerus anguipes, and the snail Lacuna variegata. Much of the grazing was on periphytic diatoms, and the tissue consumption mostly occurs during the period of slow growth, before the annual dieback, and does not affect the seasonal abundance or dominance of the plant (Norton and Benson 1983). There were fewer Green Sea Urchins (Strongylocentrotus droebachiensis) at invaded sites, apparently because they found S. muticum unpalatable (Britton-Simmons 2004). However, the snail Lacuna vincta was 2-9X more abundant on S. muticum than on native algae. This preference seems to have been acquired in the last 30 years (Britton-Simon et al. 2011).
NEA-VNoneEcological ImpactCompetition
In a lower intertidal region on the Bay of Biscay, northern Spain had negative impacts of the native red alga Gelidium spinosum, probably due to competition for light (Sanchez et al. 2005). Addition of nutrients in tide pools favored rapid growth and dominance of S. muticum, but colonization was resisted in plots with a dense canopy of the native Bifurcaria bifurcata (Sanchez and Fernandez 2005). Experimental removal of S. muticum found only limited impacts on total numbers of algal species, somewhat reducing the abundance of filamentous and foliose algae (Olabarrio et al. 2009b), or having no detectable impact on other algae (Sanchez and Fernandez 2005). High abundance of S. muticum in tide pools, in northern Portugal, was correlated with decreased abundance of native algae (Viejo et al. 1997). Modeling, based on field observations suggested that the most important feature favoring S. muticum over the native Cytoseira humilis was the persistence of non-fertile fronds of S. muticum, after reproduction, denying the space to Cytoseira humilis (Engelen and Santos 2009). Experiments with the effects of nutrient inputs found that S muticum had a complex response, and was favored by high inputs, with low variability, but not by low, highly variable inputs (Incera et al. 2009). In another set of experiments S. muticum became very abundant in tide pools with high nutrient input and mechanical disturbance (scraping with a chisel) (Bertocci et al. 2014). Nutrient fertilization of tidepools promoted the establishment and functional impacts (increased productivity and respiration) by Grateloupia turuturu and Sargassum muticum (Vieira et al. 2017).
NEA-IVNoneEcological ImpactCompetition
Sargassum muticum became the most abundant species in the intertidal of the French Atlantic coast, and was associated with a decrease in a native kelp (Laminaria digitata) (Belsher and Pommelec 1988; Cosson 1999, cited by Schaffelke and Hewitt 2007; Stiger-Pouvreau and Thouzeau 2015). It was also reported to colonize and displace Eelgrass (Zostera marina), in areas where the Eelgrass has declined, due to disturbance (Den Hartog 1997).
NEP-VIPt. Conception to Southern Baja CaliforniaEcological ImpactCompetition
Sargassum muticum colonized Giant Kelp (Macrocystis pyrifera) beds on Bird Rock, off Catalina Island, after a die-off, possibly caused by high water temperatures during El Nino. Shading by Sargassum muticum probably inhibited recolonization by the kelp. After experimental removal, the kelp recolonized the cleared areas (Ambrose and Nelson 1982). Several years later, Giant Kelp did recolonize Bird Rock (Foster & Schiel 1992, cited by Engelen et al. 2015). Removal experiments in tidepools at Little Corona del Mar in Newport Beach in southern California, showed little effect level, pool temperature, seaweed biomass and community composition, or faunal composition. Recovery of S. muticum populations was rapid (Smith 2016).
NEA-VNoneEcological ImpactFood/Prey
Experiments with a range of grazing animals, the snails Littorina littorea, L. obtusata, Gibbula, spp., and Peringia ulvae, the sea-slug Aplysia punctata, the amphipod Gammarus insensibilis, and the isopod Stenosoma nadejda, generally preferred native algae to S. muticum, while the sea urchin Paracentrotus lividus showed no preference. Preferences were variable, but experiments did suggest that S. muticum was not under high pressure from grazers (Monteiro et al. 2009; Cacabelos et al. 2010; Engelen et al. 2011). Sargassum muticum, washed up on beaches, was a major food source for the amphipod Talitrus saltator and, to a less extent, for the isopod Tylos europaeus (Rossi et al. 2009; Olabarria et al. 2010). The S. muticum wrack had higher nutrient content than that of a native alga (S. muticum), but there were not consistent differences in invertebrates using the two types of wracks (Rodil et al. 2008). Overall, the invasion of S. muticum has increased the biomass, light-use efficiency, primary production, and respiration of tide pool systems in Portugal. However, this effect disappears during the seasonal die-off of this seaweed (Vaz-Pinto et al. 2014).
NEA-IIINoneEcological ImpactCompetition
In experiments using assemblages of native algae and S. muticum reared in containers in the intertidal of Lough Hyne, Ireland, S. muticum had a negative impacts on the biomass of Fucus vesiculosus and, to a lesser extent, on Cladostephus spongiosus. In the Salcombe River estuary, England, S. muticum was able to attach to soft substrate within Eelgrass (Zostera marina) beds, because of the increased stability and decreased water movement (Tweedley et al. 2008).
NEA-IIINoneEcological ImpactHabitat Change
In experiments in Lough Hyne, Ireland, benthic animal diversity and species richness was higher in assemblages of native algae than those containing S. muticum, probably because S. muticum contains less habitat cover (Salvaterra et al. 2013). On the West Coast of Ireland, S. muticum supports dense growths of the filamentous brown alga Pylaiella littoralis, especially in sheltered sites, where the epiphyte growth, inhibited photosynthesis, growth, and caused mortality. Growth of other epiphytes and survival of S. muticum was better in more exposed sites (Baer and Stengel 2014). In the Salcombe-Knightsbridge Estuary (English Channel), Devon, England, invasion by S. muticum resulted in shorter blade length of Eelgrass (Zostera marina), and altered epibiota communities. Although S. muticum provides a more structurally complex habitat, and supports larger abundances of some taxa, its seasonal die-offs may limit the establishment of populations (DeAmicis et al. 2015).
NEA-IVNoneEcological ImpactHabitat Change
In turbid waters, Sargassum muticum replaces kelps, but provides habitat for fishes, crustaceans, and cuttlefish (Stiger-Pouvreau and Thouzeau 2015; Roux et al. 2021). Sargassum muticum does not directly compete with Eelgrass (Zostera marina, but when Eelgrass beds near Roscoff, France, are destroyed by natural shifts in the sediment, S.muticum quickly colonizes the empty spaces (den Hartog 1997).
NEA-IVNoneEconomic ImpactFisheries
Sargassum muticum can interfere with shellfishing and shellfish aquaculture, by covering the bottom, fouling shells, and equipment (Belsher and Pommelec1988); Stiger-Pouvreau and Thouzeau 2015)
NEA-VNoneEcological ImpactHabitat Change
In turbid waters, Sargassum muticum replaces kelps, but provides habitat for fishes, crustaceans, and cuttlefish (Stiger-Pouvreau and Thouzeau 2015). In a study of epifaunal invertebrates, in intertidal communities in Galicia, northern Spain, results showed that S. muticum supported levels of abundance and diversity comparable to those of two native seaweeds (Gestoso et al. 2012). At several locations on the coast of Portugal, the epifauna of S. muticum differs from that of Cystoseira humilis in composition or abundance, but not in any consistent way (Viejo et al. 1999; Engelen et al. 2013). It is used by fishes, crustaceans, and cuttlefish (Stiger-Pouvreau and Thouzeau 2015).
NEA-VNoneEconomic ImpactFisheries
Sargassum muticum can interfere with shellfishing and shellfish aquaculture, by covering the bottom, fouling shells, and equipment (Stiger-Pouvreau and Thouzeau 2015).
NEA-IVNoneEconomic ImpactShipping/Boating
Sargassum muticum entangled propellers and hampered navigation in Saint-Malo, Saint-Guénolé, and the Gulf of Morbihan (1982-1986, Belsher and Pommelec1988)
B-INoneEcological ImpactHabitat Change
Sargassum muticum supports an epiphyte community of 82 species in the Oslofjord, Norway, more than that of two major native structural plants, Fucus serratus and Zostera marina (Bjærke and Fredriksen 2005).
NEA-IINoneEconomic ImpactShipping/Boating
Dense beds of Sargassum muticum were reported to interfere with the movement of small boats and to clog their intake pipes (Critchley et al. 1986).
NEA-IINoneEconomic ImpactIndustry
Sargassum muticum was reported to clog the intakes of power plants in England (Critchley et al. 1986).
NEA-IINoneEconomic ImpactFisheries
Sargassum muticum is reported to foul fishing lines and nets, and has also interfered with oyster culture and harvesting in England and France (Critchley et al. 1986).
NEA-IINoneEconomic ImpactAesthetic
Large amounts of Sargassum muticum washing up on beaches, were an unpleasant feature on recreational beaches. Dense beds of Sargassum, were reported to interfere with swimming and recreational sailing. Sargassum muticum was also considered a threat to native marine biota and their habitats. A removal program was studied and organized by a coalition of local, regional, and national environmental agencies, in southern England (Critchley et al. 1986). In 1974-1975, a large campaign of hand-picking by volunteers was conducted, with about ~800 collecting trips and 31 metric tons collected. However, hand-collecting overlooked germlings and small plants. Mechanical removal, herbicides, and release of natural herbivores all proved to be ineffective. Specialized cutting machines, trawls, and suction devices were developed and tested, and found to be more effective, but would require continual annual use (Critchley et al. 1986).
NEP-IIIAlaskan panhandle to N. of Puget SoundEcological ImpactHabitat Change
In the San Juan Islands, Washington, Sargassum muticum supported a total of 107 epifaunal taxa, and on average supported 20 species per plant, compared to 10 species per plant on the native kelp Laminaria saccharina. Epifaunal diversity increased in area invaded by S. muticum (Giver 1999).
P292_CDA_P292 (San Juan Islands)Ecological ImpactHabitat Change
In the San Juan Islands, Washington, Sargassum muticum supported a total of 107 epifaunal taxa, and on average supported 20 species per plant, compared to 10 species per plant on the native kelp Laminaria saccharina. Epifaunal diversity increased in area invaded by S. muticum (Giver 1999).
NEA-IINoneEcological ImpactHabitat Change
On the Isle of Cumbrae, Scotland, canopies of Sargassum muticum support a higher abundance, but lower diversity of epifauna than the native Dictyota dichotoma, probably due to a more complex structure (Harries et al. 2007). In Strangford Lough, Northern Ireland, S. muticum has colonized large areas of bare substrate by 'stone-walking', attached to stones and shells, and moving by water motion. Invertebrate communities were altered under the canopies, and were dominated by smaller, more opportunistic species than the bare substrate (Strong et al. 2011). Sargassum muticum in this estuary was more heavily colonized by epiphytes and herbivorous amphipods than the natives, and appeared not to benefit from 'invader release' (Strong et al. 2009). In the Limfjorden, Denmark, S. muticum supports a similar epifaunal community in species composition to the native Halidrys siliqua, but supports a much higher density of fauna, especially amphipods, because of the greater size and complexity of S. muticum's thallus (Wernberg et al. 2005).
B-IINoneEconomic ImpactIndustry
Sargassum muticum clogged intakes of the Ringhals nuclear power plant in the province of Halland (Josefsson and Jansson 2009).
B-IINoneEconomic ImpactFisheries
In Sweden, catches of eel (Anguilla anguilla) may have been negatively influenced in some areas (Koster archipelago) through interference with fishing gear (Karlsson et al. 1995).
NEA-IINoneEcological ImpactFood/Prey
Sargassum muticum had faster rates of growth and decomposition than the native brown seaweed Halidrys siliquosa in the Limfjord, Denmark, resulting in higher productivity, and faster turnover of organic matter. Sargassum muticum was preferred to H. siliquosa by the major grazer, urchin Psammechinus miliaris, but grazing losses of S. muticum were small, compared to those due to decomposition (Pedersen et al. 2005; Pedersen et al. 2016). In feeding experiments in Germany, using the snail Littorina littorea, the urchin Psammechinus miliaris, and the isopod Idotea baltica, the native Fucus vesiculosus was preferred to S. muticum from Germany, but S. muticum and S. horneri, collected in Japan, were both more strongly avoided, suggesting that European populations have reduced chemical defenses (Schwartz et al. 2016).

Flat Periwinkles (Littorina obtusata and L. fabalis) from English Channel sites first colonized 6-40 years ago (1970s), fed as readily on S. muticum as on native Ascophyllum nodusum, in comparison to snails from later invaded areas. This difference is suggestive of behavioral or evulutionary adaptation (Kurr and Davies 2018).

In Strangford Lough, Northern Ireland, S. muticum was more densely inhabited by the amphipod Dexamine spinosa than native algae (Saccharina latissima, H. siliquosa, Fucus serratus) and more heavily grazed by the amphipod (Strong et al. 2009). In the Wadden Sea, Germany, increased abundance of the native Snake Pipefish (Entelurus aequoreus) was promoted by dense growths of S. muticum, which also supported high densities of harpacticoid copepods, food for the pipefish (Polte and Buschbaum 2008).
P058_CDA_P058 (San Pedro Channel Islands)Ecological ImpactCompetition
Sargassum muticum colonized Giant Kelp (Macrocystis pyrifera) beds on Bird Rock, off Catalina Island, after a die-off, possibly caused by high water temperatures during El Nino. Shading by Sargassum muticum probably inhibited recolonization by the kelp. After experimental removal, the kelp recolonized the cleared areas (Ambrose and Nelson 1982). Several years later, Giant Kelp did recolonize Bird Rock (Foster & Schiel 1992, cited by Engelen et al. 2015).
B-INoneEcological ImpactCompetition
In single-species experiments with Sargassum muticum and 5 native seaweed species, S. muticum grew fastest at the high temperature, 17ºC compared to the natives, but had much slower growth at 7°C. Growth of S. muticum did not differ greatly between high and low nutrient levels (Steen and Rueness 2004).
P040Newport BayEcological ImpactCompetition
Removal experiments in tidepools at Little Corona del Mar in Newport Beach in southern California, showed little effect level, pool temperature, seaweed biomass and community composition, or faunal composition. Recovery of S. muticum populations was rapid (Smith 2016).
NEA-IIINoneEcological ImpactHerbivory
Flat Periwinkles (Littorina obtusata and L. fabalis) from English Channel sites first colonized 6-40 years ago (1970s), fed as readily on S. muticum as on native Ascophyllum nodusum, in comparison to snails from later invaded areas. This could represent behavoral or evoultionary adaptation (Kurr and Davies 2018).
P090San Francisco BayEcological ImpactHabitat Change
In field cage experiments, juvenile Chinook Salmon (Oncorhynchus tshawystcha) were reared in several habitats, bare sand, Eelgrass (Zostera marina, Sargassum muticum and mixed Eelgrass-Sargassum. Growth was best in the mixed habitat, suggesting that habitat variety is important (Hughes et al. 2020).
NEP-VNorthern California to Mid Channel IslandsEcological ImpactHabitat Change
In field cage experiments, juvenile Chinook Salmon (Oncorhynchus tshawystcha) were reared in several habitats, bare sand, Eelgrass (Zostera marina, Sargassum muticum and mixed Eelgrass-Sargassum. Growth was best in the mixed habitat, suggesting that habitat variety is important (Hughes et al. 2020).
WAWashingtonEcological ImpactCompetition
The invasion of Sargassum muticum resulted in displacement of native algae, in the San Juan Islands, northern Puget Sound. A removal experiment resulted in recovery of native kelps (Britton-Simmons 2004). Modeling and experiments indicated that S. muticum invasions required a combination of disturbance and high propagule pressure (Britton-Simmons et al. 2008).
WAWashingtonEcological ImpactFood/Prey
Sargassum muticum was grazed by a high abundance, but low diversity of grazers, compared to native seaweeds. Common grazers were the amphipods Peramphithoe mea, Aoroides columbiae, Caprella laeviscula and Ischyocerus anguipes, and the snail Lacuna variegata. Much of the grazing was on periphytic diatoms, and the tissue consumption mostly occurs during the period of slow growth, before the annual dieback, and does not affect the seasonal abundance or dominance of the plant (Norton and Benson 1983). There were fewer Green Sea Urchins (Strongylocentrotus droebachiensis) at invaded sites, apparently because they found S. muticum unpalatable (Britton-Simmons 2004). However, the snail Lacuna vincta was 2-9X more abundant on S. muticum than on native algae. This preference seems to have been acquired in the last 30 years (Britton-Simon et al. 2011).
WAWashingtonEcological ImpactHabitat Change
In the San Juan Islands, Washington, Sargassum muticum supported a total of 107 epifaunal taxa, and on average supported 20 species per plant, compared to 10 species per plant on the native kelp Laminaria saccharina. Epifaunal diversity increased in area invaded by S. muticum (Giver 1999).
CACaliforniaEcological ImpactCompetition
Sargassum muticum colonized Giant Kelp (Macrocystis pyrifera) beds on Bird Rock, off Catalina Island, after a die-off, possibly caused by high water temperatures during El Nino. Shading by Sargassum muticum probably inhibited recolonization by the kelp. After experimental removal, the kelp recolonized the cleared areas (Ambrose and Nelson 1982). Several years later, Giant Kelp did recolonize Bird Rock (Foster & Schiel 1992, cited by Engelen et al. 2015)., Removal experiments in tidepools at Little Corona del Mar in Newport Beach in southern California, showed little effect level, pool temperature, seaweed biomass and community composition, or faunal composition. Recovery of S. muticum populations was rapid (Smith 2016).
CACaliforniaEcological ImpactHabitat Change
In field cage experiments, juvenile Chinook Salmon (Oncorhynchus tshawystcha) were reared in several habitats, bare sand, Eelgrass (Zostera marina, Sargassum muticum and mixed Eelgrass-Sargassum. Growth was best in the mixed habitat, suggesting that habitat variety is important (Hughes et al. 2020)., In field cage experiments, juvenile Chinook Salmon (Oncorhynchus tshawystcha) were reared in several habitats, bare sand, Eelgrass (Zostera marina, Sargassum muticum and mixed Eelgrass-Sargassum. Growth was best in the mixed habitat, suggesting that habitat variety is important (Hughes et al. 2020).

Regional Distribution Map

Bioregion Region Name Year Invasion Status Population Status
NWP-3b None 0 Native Estab
NWP-4b None 0 Native Estab
NWP-4a None 0 Native Estab
NEP-III Alaskan panhandle to N. of Puget Sound 1944 Def Estab
NEP-IV Puget Sound to Northern California 1947 Def Estab
NEP-V Northern California to Mid Channel Islands 1963 Def Estab
NEP-VI Pt. Conception to Southern Baja California 1959 Def Estab
NEA-II None 1971 Def Estab
NEA-IV None 1978 Def Estab
NEA-III None 1976 Def Estab
B-I None 1988 Def Estab
B-II None 1993 Def Estab
NEA-V None 1985 Def Estab
MED-II None 1980 Def Estab
MED-VII None 1992 Def Estab
NWP-3a None 0 Native Estab
P050 San Pedro Bay 1973 Def Estab
P170 Coos Bay 1947 Def Estab
P270 Willapa Bay 1953 Def Estab
P020 San Diego Bay 1969 Def Estab
P130 Humboldt Bay 1965 Def Estab
P030 Mission Bay 1959 Def Estab
P022 _CDA_P022 (San Diego) 1970 Def Estab
P058 _CDA_P058 (San Pedro Channel Islands) 1970 Def Estab
P027 _CDA_P027 (Aliso-San Onofre) 1971 Def Estab
P060 Santa Monica Bay 2004 Def Estab
P065 _CDA_P065 (Santa Barbara Channel) 1977 Def Estab
P069 _CDA_P069 (Central Coastal) 1973 Def Estab
P076 _CDA_P076 (Carmel) 1977 Def Estab
P080 Monterey Bay 1977 Def Estab
P090 San Francisco Bay 1973 Def Estab
P110 Tomales Bay 1973 Def Estab
P143 _CDA_P143 (Smith) 1963 Def Estab
P165 _CDA_P165 (Coos) 1978 Def Estab
P210 Yaquina Bay 1966 Def Estab
P215 _CDA_P215 (Siltez-Yaquina) 0 Def Estab
P230 Netarts Bay 1976 Def Estab
P284 _CDA_P284 (Hoh-Quillayute) 2001 Def Estab
P240 Tillamook Bay 0 Def Estab
P286 _CDA_P286 (Crescent-Hoko) 1952 Def Estab
P288 _CDA_P288 (Dungeness-Elwha) 1950 Def Estab
P290 Puget Sound 1950 Def Estab
P293 _CDA_P293 (Strait of Georgia) 0 Def Estab
P297 _CDA_P297 (Strait of Georgia) 0 Def Estab
P294 _CDA_P294 (Nooksack) 0 Def Estab
P292 _CDA_P292 (San Juan Islands) 1948 Def Estab
NWP-5 None 0 Native Estab
SP-XXI None 2000 Def Unk
P093 _CDA_P093 (San Pablo Bay) 1973 Def Estab
P056 _CDA_P056 (Los Angeles) 2004 Def Estab
WA-I None 2012 Def Estab
AR-V None 1989 Def Estab
P062 _CDA_P062 (Calleguas) 2011 Def Estab
P040 Newport Bay 2011 Def Estab
P023 _CDA_P023 (San Louis Rey-Escondido) 2011 Def Estab
MED-III None 1992 Def Unk

Occurrence Map

OCC_ID Author Year Date Locality Status Latitude Longitude

References

Mabey, Abigail L.; Catford, Jane A.; Rius.; Foggo, Andrew ; Smale, Dan A. (2022) Herbivory and functional traits suggest that enemy release is not an important mechanism driving invasion success of brown seaweeds, Biological Invasions Published online: Published online
https://doi.org/10.1007/s10530-022-02894-4

Abbott, Isabella A.; Hollenberg, George J. (1976) <missing title>, Stanford University Press, Stanford CA. Pp. <missing location>

Aguilar-Rosas, Luis Ernesto; Pedroche, Francisco Flores; Zertuche-González, José Antonio (2014) [Aquatic Invasive Species in Mexico], Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, <missing place>. Pp. 211-222

Aguilar-Rosas, Raul; Galindo, Alberto Machado (1990) Ecological aspects of Sargassum muticum in Baja California: Reproductive phenology and epiphytes, Hydrobiologia 204/205: 185-190

Ambrose, R. F.; Nelson, Bobette V. (1982) Inhibition of Giant Kelp recruitment by an introduced brown alga., Botanica Marina 25: 265-267

Andrew, N. L.; Viejo, R. M. (1998) Ecological limits to the invasion of Sargassum muticum in northern Spain., Aquatic Botany 60: 251-263

Araujo, Rita and 7 authors (2009) Checklist of benthic marine algae and cyanobacteria of northern Portugal, Botanica Marina 52(2): 24-46

Arenas, F. and 13 authors. (2006) Alien species and other notable records from a rapid assessment survey of marinas on the south coast of England., Journal of the Marine Biological Association of the United Kingdom 86: 329-1337

Arenas, F.; Fernadez, C.; Rico, J. M.; Fernandez, E,; Haya, D. (1995) Growth and reproductive strategies of Sargassum muticum (Yeno) Fensholt and Cytoseira nodicaulis (Whit.) Roberts, Scientia Marina 59(Supp. 1): 1-8

Arenas, Francisco; Fernandez, Consolacion (2002) Size-structure and dynamics in a population of Sargassum muticum (Phaeophyta), Journal of Phycology 36: 1012-1020

Arenas, Francisco; Viejo, Rosa M.; Fernández, Consolación (2002) Density-dependent regulation in an invasive seaweed: responses at plant and modular levels, Journal of Ecology 90: 820-829

Ashton, Gail ; Boos, Karin; Shucksmith, Richard; Cook, Elizabeth (2006) Rapid assessment of the distribution of marine non-native species in marinas in Scotland, Aquatic Invasions 1(4): 209-213

Augyte, Simona; Shaughnessy, Frank J. (2014) A floristic analysis of the marine algae and seagrasses between Cape Mendocino, California and Cape Blanco, Oregon, USA, Botanica Marina 57(4): 251-263

Baer, Julia; Stengel, Dagmar B. (2010) Variability in growth, development and reproduction of the non-native seaweed Sargassum muticum (Phaeophyceae) on the Irish west coast, Estuarine, Coastal and Shelf Science 90: 185-194

Baer, Julia; Stengel, Dagmar B. (2014) Can native epiphytes affect establishment success of the alien seaweed Sargassum muticum (Phaeophyceae)?, Proceedings of the Royal Irish Academy 114B(1): 41-52

Baldwin, Andy; Leason, Diane (2016) Potential Ecological impacts of Emerald Ash Borer on Maryland's Eastern Shore, In: None(Eds.) None. , <missing place>. Pp. <missing location>

Beleem, Imtiyaz B.; Kotta, Jonne; F Barboza, rancisco R. (2023) Effects of an Invasive Mud Crab on a Macroalgae-Dominated Habitat of the Baltic Sea under Different Temperature Regimes, Diversity 15(644): Published online
https:// doi.org/10.3390/d15050644

Belsher, T.; Pommellec, S. (1988) [Expansion of an alga of Japanese origin, Sargssum muticum (Yeno) Fensholt on the French coast from 1983 ro 1987], Cahiers de Biologie Marine 29: 221-231

Benali, Myriam; Djebri, Ilhem; Bellouis, Dallal; Sellam, Louiza-Nesrine;Rebzani-Zahaf, Chafika (2023) First record of drifting Sargassum muticum (Yendo) Fensholt thalli on the Algerian coasts of Cherchell and Sidi Fredj , BioInvasions Records 8(1): 575-581
https://doi.org/10.3391/bir.2019.8.3.13

Bernardi, Giacomo; Cohn, Francesca; Dominguez‑Dominguez, Omar; Kingon, Kelly; Tornabene, Luke; Robertson, D. Ross (2024) Establishment genomics of the Indo‑Pacific damselfish Neopomacentrus cyanomos, in the Greater Caribbean, Biological Invasions <missing volume>: Published online
https://doi.org/10.1007/s10530-023-03226-w

Bertocci, Iacopo and 9 authors (2014) The regime of climate-related disturbance and nutrient enrichment modulate macroalgal invasions in rockpools, Biological Invasions Published online: <missing location>

Bjærke, Marit Ruge; Rueness, Jan (2004) Effects of temperature and salinity on growth, reproduction and survival in the introduced red alga Heterosiphonia japonica (Ceramiales, Rhodophyta)., Botanica Marina 47: 373-380

Bjærke, Marit Ruge; Fredriksen, Stein (2005) Epiphytic macroalgae on the introduced brown seaweed Sargassum muticum (Yendo) Fensholt (Phaeophyceae) in Norway, Sarsia 88(5): 353-364

Boaden, P. J. S. (1995) Adventive seaweed Sargassum muticum (Yendo) Fensholt in Strangford Lough, Northern Ireland, Irish Naturalists' Journal 25(3): 111-113

Boalch, G. T.; Potts, G. W. (1977) The first occurrence of Sargassum muticum (Yendo) Fensholt in the Plymouth area, Journal of the Marine Biological Association of the United Kingdom 57: 29-31

Bold, Harold C.; Wynne, Michael J. (1978) Introduction to the Algae: Structure and Reproduction, Prentice-Hall, Englewood Cliffs, NJ. Pp. <missing location>

Boudouresque, Charles F.; Verlaque, Marc (2002) Biological pollution in the Mediterranean Sea: invasive versus introduced macrophytes., Marine Pollution Bulletin 44: 32-38

Boyd, Milton J.; Mulligan, Tim J; Shaughnessy, Frank J. (2002) <missing title>, California Department of Fish and Game, Sacramento. Pp. 1-118

Britton-Simmons, Kevin H. (2004) Direct and indirect effects of the introduced alga Sargassum muticum on benthic, subtidal communities of Washington State, USA., Marine Ecology Progress Series 277: <missing location>

Britton-Simmons, Kevin H. (2006) Functional group diversity, resource preemption and the genesis of invasion resistance in a community of marine algae, Oikos 133: 395-401

Britton-Simmons, Kevin H.; Abbott, Karen C (2008) Short- and long-term effects of disturbance and propagule pressure on a biological invasion., Journal of Ecology 96: 68-77

Britton-Simmons, Kevin H.; Pister, Benjamin; Sanchez, Inigo; Okamoto, Daniel (2011) Response of a native, herbivorous snail to the introduced seaweed Sargassum muticum, Hydrobiologia 661: 187-196

Burfeind, Dana D.; Pitt, Kylie A.; Connolly, Rod M.; Byers, James E. (2012) Performance of non-native species within marine reserves, Biological Invasions published online: <missing location>

Buschbaum, Christian ; Chapman, Amelise S.; Bettina, Saier (2006) How an introduced seaweed can affect epibiota diversity in different coastal systems., Marine Biology <missing volume>: <missing location>

Buschbaum, Christian; Lackschewitz, Dagmar; Reise, Karsten (2012) Nonnative macrobenthos in the Wadden Sea ecosystem, Journal of Ocean Management 68: 89-101

Cacabelos, Eva; Olabarria, Celia; Incera, Mónica; Troncoso, Jesús S. (2010) Do grazers prefer invasive seaweeds?, Journal of Experimental Marine Biology and Ecology 393: 182-187

California Department of Fish and Wildlife (2014) Introduced Aquatic Species in California Bays and Harbors, 2011 Survey, California Department of Fish and Wildlife, Sacramento CA. Pp. 1-36

Callahan, Mary (12/3/2023) Sea anemone native to Southern Hemiphere found in Tomales Bay, Press-Democrat <missing volume>: Published online

Carlton, J., Haugen, C., Pearse, J., Silberstien, M., Slattery, P. 1996 Exotic Species of the Monterey Bay National Marine Sanctuary. <missing URL>



Carlton, James T. (1989) <missing title>, <missing publisher>, <missing place>. Pp. <missing location>

Carlton, James T.; Eldredge, Lucius (2009) Marine bioinvasions of Hawaii: The introduced and cryptogenic marine and estuarine animals and plants of the Hawaiian archipelago., Bishop Museum Bulletin in Cultural and Environmental Studies 4: 1-202

Carreira-Flores, Diego; Rubal, Marcos; Moreira, Juan; Guerrero-Meseguer, Laura; Gomes, Pedro T.; Veiga, Puri (2023) Recent changes on the abundance and distribution of non-indigenous macroalgae along the southwest coast of the Bay of Biscay, Aquatic Botany 189(103685): Published online
https://doi.org/10.1016/j.aquabot.2023.103685

Chainho, Paula and 20 additional authors (2015) Non-indigenous species in Portuguese coastal areas, lagoons, estuaries, and islands, Estuarine, Coastal and Shelf Science <missing volume>: <missing location>

Cheang, Chi Chiu and 9 authors (2010) Low genetic variability of Sargassum muticum (Phaeophyceae) revealed by a global analysis of native and introduced populations, Journal of Phycology 46: 1063-1074

Cho, Sung Mi; Lee, Sang Mook; Ko, Yong Deok; Mattio, Lydiane; Boo, Sung Min (2012) Molecular systematic reassessment of Sargassum (Fucales, Phaeophyceae) in Korea using four gene regions, Botanica Marina 55(5): 473-484

Cohen, A.N.; Carlton, J.T.; Fountain, M.C. (1995) Introduction, dispersal and potential impacts of the green crab Carcinus maenas in San Francisco Bay, California., Marine Biology 122: 225-237

Cohen, Andrew N. 2005-2024 Exotics Guide- Non-native species of the North American Pacific Coat. https://www.exoticsguide.org/



Cohen, Andrew N. and 10 authors (2005) <missing title>, San Francisco Estuary Institute, Oakland CA. Pp. <missing location>

Cohen, Andrew N. and 12 authors (2002) Project report for the Southern California exotics expedition 2000: a rapid assessment survey of exotic species in sheltered coastal waters., In: (Eds.) . , Sacramento CA. Pp. 1-23

Cohen, Andrew N. and 22 authors (2001) <missing title>, Washington State Department of Natural Resources, Olympia. Pp. <missing location>

Cohen, Andrew N.; Carlton, James T. (1995) Nonindigenous aquatic species in a United States estuary: a case study of the biological invasions of the San Francisco Bay and Delta, U.S. Fish and Wildlife Service and National Sea Grant College Program (Connecticut Sea Grant), Washington DC, Silver Spring MD.. Pp. <missing location>

Critchley, A. T., Farnham, W. F., Thorp, C. H. (1997) On the co-occurrence of two exotic, invasive marine organisms: the brown seaweed Sargassum muticum (Yendo) Fensholt and the Spirobid tube worm Janua (Neodexiospira) brasiliensis (Grube), in association with the indigenous eelgrass, Zoster, South African Journal of Botany 63(6): 474-479

Critchley, A. T.; Farnham, W. F.; Morrell, S. L. (1986) An accournt of the attempted control of an introduced marine alga, Sargassum muticum in southern England, Biological Conservation 35: 313-322

Critchley, A. T.; Farnham, W. F.; Yoshida, T.; Norton, T. A. (1990) A bibliography of the invasive alga Sargassum muticum (Yendo) Fensholt (Fucales: Sagassaceae), Botanica Marina 33: 551-562

Critchley, A. T.; Nienhuis, P. H.; Verschuure, K. (1987) Presence and development of populations of the introduced brown alga Sargassum muticum in the southwest Netherlands, Hydrobiologia 151/152: 245-257

Critchley, Alan T. (1983a) Sargassum muticum: A morphological description of European material, Journal of the Marine Biological Association of the United Kingdom 63: 813-824

Critchley, Alan T. (1983b) Sargassum muticum:A taxonomic history including worldwide and Western Pacific distributions, Journal of the Marine Biological Association of the United Kingdom 63: 617-625

Critchley, Alan T., Thorp, C. H. (1985) Janua (Dexiospira) brasilliensis (Grube) (Polychaeta: Spirorbidae): a new record from the south-west Netherlands, Zoologische Bijdragen <missing volume>(31): 1-8

Curiel, D.; Bellemo, G.; Marzocchi, M.; Scattolin, M.; Parisi, G. (1998) Distribution of introduced Japanese macroalgae Undaria pinnatifida, Sargassum muticum (Phaeophyta) and Antithamnion pectinatum (Rhodophyta) in the Lagoon of Venice, Hydrobiologia 385: 17-22

Davidson, Alisha D.; Campbell, Marnie L.; Hewitt, Chad L.; Schaffelke, Britta (2015) Assessing the impacts of nonindigenous marine macroalgae: an update of current knowledge, Botanica Marina 58(2): 55-79

de Rivera, Catherine, and 27 authors (2005) Broad-scale non-indigenous species monitoring along the West Coast in National Marine Sanctuaries and National Estuarine Research Reserves report to National Fish and Wildlife Foundation, National Fish and Wildlife Foundation, Washington, D.C.. Pp. <missing location>

De Vries, Dennis R.;Wright, Russel A; De Vries, Tammy A. (2006) Daphnia lumholtzi in the Mobile River drainage, USA: Invasion of a habitat that experiences salinity, Journal of Freshwater Ecology 21(3): 527-530

DeAmicis, Stacey; Foggo, Andrew (2015) Long-term field study reveals subtle effects of the invasive alga Sargassum muticum upon the epibiota of Zostera marina, PLOS ONE 10(9): e0137861

den Hartog, C. (1997) Is Sargassum muticum a threat to eelgrass beds?, Aquatic Botany 58: 37-41

Dethier, Megan N.; Carl Schoch, G. (2005) The consequences of scale: assessing the distribution of benthic populations in a complex estuarine fjord, Estuarine, Coastal and Shelf Science 62: 253-270

Deysher, Larry; Norton, Trevor A. (1982) Dispersal and colonization in Sargassum muticum (Yendo) Fensholt, Journal of Experimental Marine Biology and Ecology 56: 179-195

Deysher, Lawrence R. (1984) Reproductive phenology of newly introduced populations of the brown alga Sargassum muticum (Yendo) Fensholt, Hydrobiologia 116/117: 403-407

Dias, P. Joana and 9 authors (2021) Multiple introductions and regional spread shape the distribution of the cryptic ascidian Didemnum perlucidum in Australia:: an important baseline for management under climate change, Aquatic Invasions 16: 297-313

Dickey, James W. E. and 9 authors (2021) Breathing space: deoxygenation of aquatic environments can drive differential ecological impacts across biological invasion stages, Biological Invasions Published online: <missing location>

Dixon, Rainbo R. M. and 6 authors (2014) North meets south;Taxonomic and biogeographic implications of a phylogenetic assessment of Sargassum subgenera Arthrophycus and Bactrophycus (Fucales, Phaeophyceae), Phycologia 53(1): 15-22

Engelen, Aschwin H.; Henriques, Nuno; Monteiro, Carla; Santos, Rui (2011) Mesograzers prefer mostly native seaweeds over the invasive brown seaweed Sargassum uticum., Hydrobiologia 669: 157-165

Engelen, Aschwin H. and 14 authors (2015) Circumglobal invasion by the brown seaweed Sargassum muticum, Oceanography and Marine Biology, an Annual Review 53: 81-126

Engelen, Aschwin H.; Espirito-SantoSimoes, Cristina; Tiago; Monteiro, Carla; Serrao, Ester A.; Pearson, Gareth A.; Santos, Rui O. P. (2008) Periodicity of propagule expulsion and settlement in the competing native and invasive brown seaweeds, Cystoseira humilis and Sargassum muticum (Phaeophyta), European Journal of Phycology 43(3): 275-282

Engelen, Aschwin H.; Primo, Ana L.; Cruz, Teresa; Santos, Rui (2013) Faunal differences between the invasive brown macroalga Sargassum muticum and competing native macroalgae, Biological Invasions 15(1): 171-183

Engelen, Aschwin; Santos, Rui (2009) Which demographic traits determine population growth in the invasive brown seaweed Sargassum muticum?, Journal of Ecology 97: 675-684

Eno, N. Clare; Clark, Robin A.; Sanderson, William G. (1997) <missing title>, Joint Nature Conservation Committee, Peterborough. Pp. <missing location>

Espinoza, J. (1990) The southern limit of Sargassum muticum (Yendo) Fensholt (Phaeophyta, Fucales) in the Mexican Pacific, Botanica Marina 33: 193-196

Farnham, W. F. (1980) Studies on aliens in the marine flora of southern England, In: Price, J. H., Irvine, D. E. G., and Farnham, W. F.(Eds.) The Shore Environment. , London. Pp. 875-914

Fernandez, C. (1999) Ecology of Sargassum muticum (Phaeophyta) on the north coast of Spain: IV. Sequence of colonization on a shore, Botanica Marina 42: 553-562

Fernández, C. (2020) Boom-bust of Sargassum muticum in northern Spain: 30 years of invasion, European Journal of Phycology 55(3): 285=-295

Fernandez, C., Gutierrez, L. M.; Rico, J. M. (1990) Ecology of Sargassum muticum on the North Coast of Spain, Botanica Marina 33: 423-428



Fletcher, R.L., Farrell, P. (1999) Introduced brown algae in the North East Atlantic, with particular respect to Undaria pinnatifida (Harvey) Suringar, Helgoländer Meeresuntersuchungen 52: 259-275

Gestoso, Ignacio; Olabarria, Celia; Troncoso, Jesus S. (2010) Variability of epifaunal assemblages associated withn native and invasive macroalgae, Marine and Freshwater Research 61: 724-731

Gestoso, Ignacio; Olabarria, Celia; Troncoso, Jesus S. (2012) Effects of macroalgal identity on epifaunal assemblages: native species versus the invasive species Sargassum muticum, Helgoland Marine Research 66: 159-166

Gittenberger, Adriaan; Rensing, Marjolein; Stegenga, Herre; Hoeksema, Bert (2010) Native and non-native species of hard substrata in the Dutch Wadden Sea, Nederlandse Faunistiche Mededelingen 33: 20-76

Giver, Karen (1999) <missing title>, Western Washington University, <missing place>. Pp. 1-193

Guiry, M. D.; Guiry, G. M. 2004-2023 AlgaeBase. https://www.algaebase.org/



Guiry, M.D. 1998-2016 Seaweed Site. <missing URL>



Gunnill, F.C. (1982) Effects of plant size and distribution on the numbers of invertebrate species and individuals inhabiting the brown alga Pelvetia fastigiata, Marine Biology 69: 263-280

Gunnill, Frederic C. (1985) Population fluctuations of seven macroalgae in southern California during 1981-1982, including effects of severe storms and an El Nino, Journal of Experimental Marine Biology and Ecology 85: 149-164

Hales, J. M.; Fletcher, R. L. (1989) Studies on the recently introduced brown alga Sargassum muticum. IV. The effect of temperature, salinity, and irradiance on germling growth, Botanica Marina 32: 167-176

Harries, D. B.; Harrow, S.; Wilson, J. R.; Mair, J. M.; Donnan, D.W. (2007) The establishment of the invasive alga Sargassum muticum on the west coast of Scotland: a preliminary assessment of community effects, Journal of the Marine Biological Association of the United Kingdom 87: 1057-1067

Hopkins, C.C.E. (2002) Invasive aquatic species of Europe: Distribution, impacts, and management, Kluwer Academic Publishers, <missing place>. Pp. 240-253

Incera, Monica; Olabarria, Celia; Cacabelos, Eva; Cesar, Javier; Troncoso, Jesus S. (2010) Distribution of Sargassum muticum on the North West coast of Spain: Relationships with urbanization and community diversity, Continental Shelf Research published online: <missing location>

Incera, Mónica; Olabarria, Celia; Troncoso, Jesús S.; López, Jesús (2007) Response of the invader Sargassum muticum to variability in nutrient supply, Marine Ecology Progress Series 377: 91-101,

Inderjit; Chapman, David ; Ranelletti, Marla ; Kaushik, Shalini (2006) Invasive marine algae: an ecological perspective., Botanical Review 72(2): 153-178

Johnston, Matthew W.; Akins, John L. (2016) The non‑native royal damsel (Neopomacentrus cyanomos) in the southern Gulf of Mexico: An invasion risk?, Marine Biology 163(12): Published online
DOI 10.1007/s00227-015-2777-7

Josefsson, Melanie; Jansson, Kristina 2009 NOBANIS: Invasive Alien Species Fact Sheet: <em>Sargassum muticum</em>. <missing URL>



Josselyn, Michael N.; West, John A. (1985) The distribution and temporal dynamics of the estuarine macroalgal community of San Francisco Bay, Hydrobiologia 129: 132-152

Kaplanis, Nikolas John; Harris, Jill L.; Smith, Jennifer E. (2016) Distribution patterns of the non-native seaweeds Sargassum horneri(Turner) C. Agardh and Undaria pinnatifida (Harvey) Suringar on the San Diego and Pacific coast of North America, Aquatic Invasions 11: In press

Karlsson, J.; Loo, L-O (1999) On the distribution and the continuous expansion of the Japanese seaweed- Sargassum muticum- in Sweden., Botanica Marina 42(3): 285-294

Karsiotis, Susanne I.; , Pierce, Lindsey R.; Brown, Joshua E.; Stepien, Carol A. (2012) Salinity tolerance of the invasive round goby: Experimental implications for seawater ballast exchange and spread to North American estuaries, Journal of Great Lakes Research 38: 121-128

Kerrison, Philip; Le, Hau, Nhu (2016) Environmental factors on egg liberation and germling production of Sargassum muticum, Journal of Applied Phycology 28: 481-489

Knoepffler-Peguy, Michele; Belsher, Thomas; Boudouresque, Charles F. (1985) Sargassum muticum begins to invade the Mediterranean., Aquatic Botany 23: 291-295

Kraan, Stefan (2008) Sargassum muticum (Yendo) Fensholt in Ireland: an invasive species on the move, Journal of Applied Phycology 20: 825-832

Lang, Anne C.; Buschbaum, Christian (2010) Facilitative effects of introduced Pacific oysters on native macroalgae are limited by a secondary invader, the seaweed Sargassum muticum, Journal of Sea Research 63: 119-128

Leidenberger, Sonja; Obst, Matthias; Kulawik, Robert; Stelzer, Kerstin; Heyer, Karin; Hardisty, Alex; Bourlat, Sarah J. (2015) Evaluating the potential of ecological niche modelling as a component in marine non-indigenous species risk assessments, Marine Pollution Bulletin 97: 470-487

Liu, Wenliang; Liang, Xiaoli ; Zhu, Xiaojing (2015) A new record and mitochondrial identification of Synidotea laticauda Benedict, 1897 (Crustacea: Isopoda: Valvifera: Idoteidae) from the Yangtze Estuary, China, Zootaxa 4294: 371-380

Locke, Andrea; Hanson, John Mark (2009) Rapid response to non-indigenous species. 1. Goals and history of rapid response in the marine environment., Aquatic Invasions 4(1): 237-247

Looby, Audrey; Ginsburg, David W. (2021) Nearshore species biodiversity of a marine protected area off Santa Catalina Island, California, Western North American Naturalist 81(1): 113-130

Lyons, Devin A.; Scheibling, Robert E. (2009) Range expansion by invasive marine algae: rates and patterns of spread at a regional scale, Diversity and Distributions 15: 762-775

Mach, Megan E.; Levings, Colin D.; Chan, Kai M. A. (2016) Nonnative species in British Columbia eelgrass beds spread via shellfish aquaculture and stay for the mild climate, Estuaries and Coasts Published online: <missing location>

Maloney, E.; Fairey, R.; Lyman, A.; Reynolds, K.; Sigala, M. (2006) <missing title>, California Department of Fish and Game, Office of Spill Prevention and Response, Sacramento. Pp. <missing location>

Milledge, John J.; Nielsen, Birthe V.; Bailey, David (2016) High-value products from macroalgae: the potential uses of the invasive brown seaweed, Sargassum muticum, Reivew of Environmental Science and Biotechnology 57: 67-88

Minchin, Dan; Cook, Elizabeth J.; Clark, Paul F. (2013) Alien species in British brackish and marine waters, Aquatic Invasions 8: in press

Mineur, Frederic ; Belsher, Thomas; Johnson, Mark P.; Maggs, Christine A.; Verlaque, Marc (2007) Experimental assessment of oyster transfers as a vector for macroalgal introductions., Biological Conservation 136: 237-247

Mineur, Frederic; Johnson, Mark P.; Maggs, Christine A.; Stegenga, Herre (2007) Hull fouling on commercial ships as a vector of macroalgal introduction., Marine Biology 151: 1299-1307

Mineur, Frederic; Johnson, Mark P.; Maggs, Christine A. (2008) Non-indigenous marine macroalgae in native communities: a case study in the British Isles., Journal of the Marine Biological Association of the United Kingdom 88(4): 693-698

Mineur, Frederic; Johnson, Mark P. ; Maggs, Christine A. (2008) Macroalgal introductions by hull fouling on recreational vessels: Seaweeds and sailors., Environmental Management 42: 667-676

Monniot, Claude (1983) Littoral ascidians of Guadeloupe II. Phlébobranches, Bulletin du Museum National d'Histoire Naturelle. 4e Serie. Section A. Zoologie, Biologie et Ecologie Animales 5: 51-71

Monteiro, Carla A.; Engelen, Aschwin H.; Santos, Rui O. P. (2009a) Macro- and mesoherbivores prefer native seaweeds over the invasive brown seaweed Sargassum muticum: a potential regulating role on invasions, Marine Biology 156: 2505-2515

Monteiro, Carla; Engelen, Aschwin H.; Serrao, Ester A.; Santos, Rui (2009b) Habitat differences in the timing of reproduction of the invasive alga Sargassum muticum (Phaeophyta, Sargassaceae) over tidal and lunar cycles, Journal of Phycology 45: 1-7

Moore, Colin G.; Harries, Dan B. (2009) Appearance of Heterosiphonia japonica (Ceramiales: Rhodophyceae) on the west coast of Scotland, with notes on Sargassum muticum (Fucales: Heterokontophyta), Marine Biodiversity Records 2: 1-5

Moyle, Peter B. (2002) Inland Fishes of California, revised and expanded, University of California Press, Berkeley CA. Pp. <missing location>

Muñoz, Ricardo González; Lauretta, Daniel; Bazterrica, María Cielo; Tapia, Francisco Alejandro Puente; Garese, Agustín; Bigatti, Gregorio E.; Penchaszadeh, Pablo E.; Lomovasky, Betina ; Acuña, Fabián H. (2023) Mitochondrial and nuclear gene sequencing confirms the presence of the invasive sea anemone Diadumene lineata (Verrill, 1869) (Cnidaria: Actiniaria) in Arg, PeerJ 11(e16479): Published online
DOI 10.7717/peerj.16479

Murray, Cathryn Clarke; Pakhomov, Evgeny A.; Therriault, Thomas W. (2011) Recreational boating: a large unregulated vector transporting marine invasive species, Diversity and Distributions 17: 1161-1172

Neushul, Michael, Amsler, Charles D., Reed, Daniel C., Lewis, Raymond J. (1989) Dispersal of marine plants for aquacultural purposes, Journal of Shellfish Research 7(3): 555

Norton, T. A.; Benson, M. R, (1983) Ecological interactions between the brown seaweed Sargassum muticum and its associated fauna, Marine Biology 75: 169-177

Norton, Trevor A. (1981) Sargassum muticum on the Pacific coast of North America, Proceedings of the International Seaweed Symposium 8: 449-456

Norton, Trevor A. (1981) The varied dispersal mechanisms of an invasive seaweed, Sargassum muticum., Phycologia 20(2): 110-111

Norton, Trevor A.; Fetter, Richard (1981) The settlement of Sargassum muticum propagules in stationary and flowing water, Journal of the Marine Biological Association <missing volume>: <missing location>

Nunez-Lopez, R. A.; Casas Valdez, M. (1998) Seasonal variation of seaweed biomass in San Ignacio Lagoon, Baja California Sur, Mexico, Botanica Marina 41: 421-426

Nyberg, Cecilia D.; Wallentinus, Inger (2005) Can species traits be used to predict marine macroalgal introductions?, Biological Invasions 7: 265-279

O'Shaughnessy, Kathryn A.; Lyons, David; Ashelby,Christopher W; R Counihan, Randall; Pears, Eliot; Taylor; Davies, Rebecca; PStebbing, aul D. (2-023) Rapid assessment of marine non-native species in Irish marinas, Management of Biological Invasions 14: 245–267
, https://doi.org/10. 3391/mbi.2023.14.2.05 Received: 4 August 2022

Occhipinti Ambrogi, Anna (2000) Biotic invasions in a Mediterranean lagoon., Biological Invasions 2: 165-176

Olabarria, Celia; Incera, Mónica; Garrido, Josefina; Rossi, Francesca (2010) The effect of wrack composition and diversity on macrofaunal assemblages in intertidal marine sediments, Journal of Experimental Marine Biology and Ecology 396: 18-26

Olabarria, Celia; Rodil, Iván F.; Incera, Mónica; Troncoso, Jesús S. (2009a) Limited impact of Sargassum muticum on native algal assemblages from rocky intertidal shores, Marine Environmental Research 67: 153-158

Olabarria, Celia; Rossi, Francesca; Rodil, Ivan F.; Quintas, Patricia; Troncoso, Jesús S. (2009b) Use of hierarchical designs to detect scales of heterogeneity in the invasive species Sargassum muticum, Scientia Marina 73(3): 507-514

Oliveira, Otto M. P. and 24 authors (2016) Census of Cnidaria (Medusozoa) and Ctenophora from South American marine waters, Zootaxa 4194: 1-256

Pedersen, Morten Foldager; Stæhr, Peter Anton; Wernberg, Thomas; Thomsen, Mads Solgaard (2005) Biomass dynamics of exotic Sargassum muticum and native Halidrys siliquosa in Limfjorden, Denmark—Implications of species replacements on turnover rates, Aquatic Botany 83: 31-47

Peña, V.; Bárbara, I.; Grall, J.; Maggs, C. A.; Hall-Spencer, J. M. (2014) The diversity of seaweeds on maerl in the NE Atlantic, Marine Biodiversity 44: 533-551

Pérez-Cirera, José Luis; Cremades, Javier; Bárbara, Ignacio (1989) Precisiones sistemáticas y sinecológicas sobre algunas algas nuevas para Galicia o para las costas Atlánticas de la península Ibérica., Anales del Jardin Botanico de Madrid 46(1): 35-45

Piazzi, Luigi; Balata, David (2009) Invasion of alien macroalgae in different Mediterranean habitats., Biological Invasions 11: 193-204

Plouguerne, Erwan; Hellio, Claire; Deslandes, Eric; Veron, Benoit; Stiger-Pouvreau, Vale rie (2008) Anti-microfouling activities in extracts of two invasive algae: Grateloupia turuturu and Sargassum muticum., Botanica Marina 51: 202-208

Plouguerne, Erwan; Le Lann, Klervi; Connan, Solene; Jechoux,Gregory; Deslandes, Eric; Stiger-Pouvreau, Valerie (2006) Spatial and seasonal variation in density, reproductive status, length and phenolic content of the invasive brown macroalga Sargassum muticum (Yendo) Fensholt along the coast of Western Brittany (France), Aquatic Botany 85: 337-344

Polte, Patrick; Buschbaum, Christian (2008) Native pipefish Entelurus aequoreus are promoted by the introduced seaweed Sargassum muticum in the northern Wadden Sea, North Sea, Aquatic Biology 3: 11-18

Pongparadon, Supattra; Zuccarello, Giuseppe C.; Prathep, Anchana (2017) High morpho-anatomical variability in Halimeda macroloba (Bryopsidales, Chlorophyta) in Thai waters, Phycological Research 65: 136-145
doi: 10.1111/pre.12172

Raoux, Aurore; Pezy, Jean-Philippe; Sporniak, Thomas; Dauvin, Jean-Claude (2021) Does the invasive macro-algae Sargassum muticum (Yendo) Fensholt, 1955 offer an appropriate temporary habitat for mobile fauna including non-indigenous species?, Ecological Indicators 126(107624): Published online

Reed, Peggy; Francis-Floyd, Ruth; Klinger, Ellen; Petty, Denise 1996 Monogenean Parasites of Fish. https://thefishsite.com/articles/monogenean-parasites-of-fish



Ribera, Maria Antonia, Boudouresque, Charles-Francois (1995) Introduced marine plants, with special reference to macroalgae: mechanisms and impact., Progress in Phycological Research 11: 187-268

Riggs, Sharon R. (2011) <missing title>, Padilla Bay NERR, Padilla Bay WA. Pp. 5

Rodil, Iván F.; Olabarria, Celia; Lastra, Mariano ; López, Jesús (2008) Differential effects of native and invasive algal wrack on macrofaunal assemblages inhabiting exposed sandy beaches., Journal of Experimental Marine Biology and Ecology 358: 1-13

Rossi, Francesca; Incera, Monica; Callier, Myriam; Olabarria, Celia (2011) Effects of detrital non-native and native macroalgae on the nitrogen and carbon cycling in intertidal sediments, Marine Biology 158: 2705-2715

Rossi, Francesca; Olabarria, Celia; Incera, Mónica; Garrido, Josefina (2009) The trophic significance of the invasive seaweed Sargassum muticum in sandy beaches, Journal of Sea Research in press: <missing location>

Rueness, Jan (1989) Sargassum muticum and other introduced Japanese macroalgae: biological pollution of European coasts, Marine Polution Bulletin 20(4): 173-176

Sabour, Brahim; Reani, Abdeltif; EL Magouri, Hachem; Haroun, Ricardo (2013) Sargassum muticum (Yendo) Fensholt (Fucales, Phaeophyta) in Morocco, an invasive marine species new to the Atlantic coast of Africa, Aquatic Invasions 8: published online

Salvaterra, Tania; Green, Dannielle S.; Crowe, Tasman P.; O’Gorman, Eoin J. (2013) Impacts of the invasive alga Sargassum muticum on ecosystem functioning and food web structure, Biological Invasions published online: <missing location>

Sánchez, Íñigo; Fernández, Consolación (2006) Resource availability and invasibility in an intertidal macroalgal assemblage., Marine Ecologiy Progress Series 313: 85-94

Sanchez, Inigo; Fernandez, Consolacion; Arrontes, Julio (2005) Long-term changes in the structure of intertidal assemblages after invasion by Sargassum muticum (Phaeophyta), Journal of Phycology 41: 942-949

Sanchez, Iniigo; Fernandez, Consolacion (2005) Impact of the invasive seaweed Sargassum muticum (Phaeophyta) on an intertidal macroalgal assemblage, Journal of Phycology 41: 923-930

Scagel, R.F. (1956) Introduction of a Japanese alga, Sargassum muticum, into the Northeast Pacific., Fisheries Research Papers, Washington Department of Fisheries 1(4): 49-58

Scagel, R.F., Gabrielson, P.W., Garbary, D.J., Golden, L., Hawkes, M.W., Lindstrom, S.C., Oliveira, J.C., Widdowson, T.B. (1989) <missing title>, University of British Columbia, Vancouver. Pp. <missing location>

Schaffelke, Britta; Hewitt, Chad L. (2007) Impacts of introduced seaweeds., Botanica Marina 50: 397-417

Schaffelke, Britta; Smith, Jennifer E.; Hewitt, Chad L. (2006) Introduced macroalgae- a growing concern, Journal of Applied Phycology 18: 529-541.

Sfriso, A.; Facca, C. (2013) Annual growth and environmental relationships of the invasive species Sargassum muticum andUndaria pinnatifida in the lagoon of Venice, Estuarine, Coastal and Shelf Science 129: 162-172

Silva, Paul C. (1979) San Francisco Bay Bay: The urbanized estuary; Investigations into the Natural History of San Francisco Bay and Delta With Reference to the Influence of Man, Pacific Division of the American Association for the Advancement of Science, San Francisco. Pp. 287-346

Sjøtun, Kjersti; Eggereide, Sarah F.; Høisæter, Tore (2007) Grazer-controlled recruitment of the introduced Sargassum muticum (Phaeophyceae, Fucales) in northern Europe., Marine Ecology Progress Series 342: 127-138

Sloan, N. A., Bartier, P. M. (2004) Introduced marine species in the Haida Gwaii region, British Columbia., Canadian Field-Naturalist 118(1): 77-84

Soors, Jan; Faasse, Marco; Stevens, Maarten; Verbessem, Ingrid; De Regge, Nico;Van den Bergh, Ericia (2010) New crustacean invaders in the Schelde estuary (Belgium), Belgian Journal of Zoology 140: 3-10

Stæhr, Peter A.; Pedersen, Morten F.; Thomsen, Mads S.; Wernberg, Thomas; Krause-Jensen, Dorte (2000) Invasion of Sargassum muticum in Limfjorden (Denmark) and its possible impact on the indigenous macroalgal community., Marine Ecology Progress Series 207: 79-88

Steen, H.; Rueness, J. (2004) Comparison of survival and growth in germlings of six fucoid species (Fucales, Phaeophyceae) at two different temperature and nutrient levels, Sarsia 89: 175-183

Steen, Henning (2004) Effects of reduced salinity on reproduction and germling development in Sargassum muticum (Phaeophyceae, Fucales), European Journal of Phycology 39: 293-299

Stiger-Pouvreau, Valérie; Thouzeau, Gérard (2015) Marine species introduced on the French Channel-Atlantic coasts: a review of main biological invasions and impacts, Open Journal of Ecology, 5: 227-257

Streftaris, N.; Zenetos, A. (2006) Alien marine species in the Mediterranean - The 100 ‘worst invasives’ and their impact., Mediterranean Marine Science 7(1): 87-118

Strong, James A.; Dring, Matthew J. (2011) Macroalgal competition and invasive success: testing competition in mixed canopies of Sargassum muticum and Saccharina latissima, Botanica Marina 54: 223-229

Strong, James A.; Dring, Matthew J.; Maggs, Christine A. (2006) Colonisation and modification of soft substratum habitats by the invasive macroalga Sargassum muticum, Marine Ecology Progress Series 321: 87-97

Strong, James A.; Maggs, Christine A.; Johnson, Mark P. (2009) The extent of grazing release from epiphytism for Sargassum muticum (Phaeophyceae) within the invaded range, Journal of the Marine Biological Association of the United Kingdom 89(2): 303-314

Thibaut, Thierry; Blanfune, Aurelie; Verlaque, Marc; Boudouresque, Charles-Francois; Ruitton. Sandrine (2015) The Sargassum conundrum: very rare, threatened or locally extinct in the NW Mediterranean and still lacking protection, Hydrobiologia Published online: <missing location>

Thomsen, Mads S. ; Wernberg, Thomas ; Tuya, Fernando; Silliman, Brian R. (2009) Evidence for impacts of nonindigenous macroalgae: a meta-analysis of experimental field studies, Journal of Phycology 35: 812-819

Thomsen, Mads S.; Wernberg, Thomas; Stæhr, Peter;Krause-Jensen, Dorte; Risgaard-Petersen, Nils; Silliman, Brian R. (2007) Alien macroalgae in Denmark - a broad-scale national perspective., The Korean Journal of Systematic Zoology 3: 61-72

Thomsen, Mads S.; Wernberg, Thomas; Stæhr, Peter A.; Pedersen, Morten F. (2006) Spatio-temporal distribution patterns of the invasive macroalga Sargassum muticum within a Danish Sargassum-bed, Helgoland Marine Research 60: 50-58

Trowbridge, Cynthia D.; Little, Colin; Pilling, Graham M.; Stirling, Penny; Miles, Alison (2011) Decadal-scale changes in the shallow subtidal benthos of an Irish marine reserve, Botanica Marina 54: 497-506

Tweedley, James R.; Jackson, Emma L.; Attrill, Martin J. (2008) Zostera marina seagrass beds enhance the attachment of the invasive alga Sargassum muticum in soft sediments, Marine Ecological Progress Series 354: 305-309

University of Alaska Southeast Herbarium Portal 2016 Catalog #: ALAJ00982, Secondary Catalog #: 6516 <em>Sargassum muticum</em>. Collector L. Barr, 03 May 1974, Southern Sea Otter Sound. <missing URL>



Valentine, Joseph P.; Magierowski, Regina H.; Johnson, Craig R. (2007) Mechanisms of invasion: establishment, spread and persistence of introduced seaweed populations., Botanica Marina 50: 351-360

Vaz-Pinto, F. and 5 authors (2013) Functional diversity and climate change: effects on the invasibility of macroalgal assemblages, Biological Invasions 15: 1833-1846

Vaz-Pinto, Fátima; Martínez, Brezo; Olabarria, Celia; Arenas, Francisco (2014) Neighbourhood competition in coexisting species: The native Cystoseira humilis vs the invasive Sargassum muticum, Journal of Experimental Marine Biology and Ecology 454: 32-41

Vaz-Pinto, Fátima; Olabarria, Celia; Arenas, Francisco (2013) Role of top-down and bottom-up forces on the invasibility of intertidal macroalgal assemblages, Journal of Sea Research 76: 178-186

Vaz-Pinto, Fátima; Olabarria, Celia; Arenas, Francisco (2014) Ecosystem functioning impacts of the invasive seaweed Sargassum muticum (fucales, phaeophyceae), Journal of Phycology 50: 108-116

Vázquez-López, Horacio; Cházaro-Olvera, Sergio; Vargas-Téllez, Irma ; Molina-Ortega, Madeline Getzemany (2017) Description of first zoeal stage of Cardisoma crassum Smith, 1870 (Crustacea: Decapoda: Gecarcinidae), Journal of Natural History 51(15-19): 837-842

Verlaque, Marc (2001) Checklist of the macroalgae of Thau Lagoon (Herault, France), a hot spot of marine species introduction in Europe, Oceanologia Acta 24(1): 29-49

Viejo, R. M.; Arrontes, J.; Andrew, N. L. (1995) An experimental evaluation of the effect of wave action on the distribution of Sargassum muticum in northern Spain, Botanica Marina 38: 437-441

Viejo, Rosa M. (1997) The effects of colonization by Sargassum muticum on tidepool macroalgal assemblages, Journal of the Marine Biological Association of the United Kingdom 77: 325-340

Viejo, Rosa M. (1999) Mobile epifauna inhabiting the invasive Sargassum muticum and two local seaweeds in northern Spain, Aquatic Botany 64: 131-149

Vye, Siobhan R.; Emmerson, Mark C.; Arenas, Francisco; Dick, Jaimie T. A.; O’Connor, Nessa E. (2015) Stressor intensity determines antagonistic interactions between species invasion and multiple stressor effects on ecosystem functioning, Oikos 124: 1005-1012

Walker, D. I., Kendrick, G. A. (1998) Threats to macroalgal diversity: Marine habitat destruction and fragmentation, pollution and introduced species, Botanica Marina 41: 105-112

Want, Andrew; Matejusova, Iveta; Kakkonen, Jenni E. (2023) The establishment of the invasive non-native macroalga Sargassum muticum in the north of Scotland, Journal of the Marine Biological Association of the United Kingdom 1-3(e69): e69
https://doi.org/ 10.1017/S0025315423000577

Wernberg, T.; Thomsen, M. S.; Stæhr, P. A.; Pedersen, M. F. (2000) Comparative phenology of Sargassum muticum and Halidrys siliquosa (Phaeophyceae: Fucales) in Limfjorden, Denmark, Botanica Marina 43: 31-39

Wernberg, Thomas; Thomsen, Mads S.; Staehr, Peter A.; Pedersen, Morten F. (2004) Epibiota communities of the introduced and indigenous macroalgal relatives Sargassum muticum and Halidrys siliquosa in Limfjorden (Denmark), Helgoland Marine Research 58: 154-161

White, Laura F.; Orr, Lindsay C. (2011) Native clams facilitate invasive species in an eelgrass bed, Marine Ecology Progress Series 424: 87-95

White, Laura F.; Shurin, Jonathan B. (2007) Diversity effects on invasion vary with life history stage in marine macroalgae., Oikos 116: 1193-1203

White, Laura F.; Shurin, Jonathan B. (2011) Density dependent effects of an exotic marine macroalga on native community diversity, Journal of Experimental Marine Biology and Ecology 405: 111-119

Williams, Susan L. (2007) Introduced species in seagrass ecosystems: Status and concerns., Journal of Experimental Marine Biology and Ecology 350: 89-110

Williams, Susan L.; Smith, Jennifer E. (2007) A global review of the distribution, taxonomy, and impacts of introduced seaweeds., Annual Review of Ecology, Evolution, and Systematics 38: 327-59