Skip to main content

REVIEW article

Front. Mar. Sci., 29 April 2022
Sec. Marine Ecosystem Ecology
Volume 9 - 2022 | https://doi.org/10.3389/fmars.2022.871204

It’s the Little Things: The Role of Microscopic Life Stages in Maintaining Kelp Populations

Matthew S. Edwards*
  • Department of Biology, San Diego State University, San Diego, CA, United States

Kelp forests are experiencing broad scale declines in abundance and shifts in latitudinal ranges in many areas of the world due to numerous environmental stressors, especially those associated with climate change. While the majority of studies on kelp ecology have focused exclusively on the macroscopic sporophyte life stage, a growing number of studies is showing quite convincingly that the microscopic zoospore and gametophyte life stages can be important to establishing spatial patterns in these marine forests, and in facilitating their recovery following deforestation. Unfortunately, their microscopic sizes have made them difficult to study in the field, leading to a ‘black box’ surrounding what we know about their ecologies. However, advances in experimental methodologies and a growing number of laboratories studying kelp microscopic life stages are revealing how they are affected by variability in environmental conditions, and are providing a clearer picture of how kelp forests might respond to a changing ocean climate. These studies have largely revealed that kelps can disperse over a wide range of distances, enhanced by the synchronous release and vertical transport of zoospores into shallower water, and by floating rafts of reproductive thalli. Settlement on the benthos is facilitated by both positive and negative chemotaxis, and by active selection of microhabitats that favor their growth and survival. Following settlement and subsequent germination, the haploid gametophytes can delay their development and form a bank of microscopic forms that persist during periods that are unfavorable for the large sporophyte populations, or they can develop and undergo sexual reproduction during which they respond to variability in environmental conditions. In particular, they are strongly affected by increases in irradiance (PAR, UVA + UVB) and temperature, decreases in nutrients and salinity, and by sedimentation and grazing. However, the manner in which they respond to these stressors varies among species and with their geographic distributions, which is integral to establishing biogeographic patterns in the large sporophyte populations. Given these factors are expected to change as the ocean climate changes, these species-specific responses have significant implications for future patterns of distribution and abundance of these iconic marine forests.

Introduction

Much has changed in our understanding of kelp microscopic life stages during the almost three decades since I first sterilized experimental reef patches in Carmel Bay, CA, USA to evaluate the role of a bank of microscopic life stages in the brown alga Desmarestia ligulata that delayed their reproduction during the winter and then produced sporophytes the following spring. Since then, the number of studies on the ecology and physiology of kelp microscopic life stages has grown tremendously, providing a much clearer picture of the role they play in kelp forest ecology and in the resiliency of these forests in the face of environmental variability. The reasons for this increase in studies are varied, but presumably include advances in or ability to study these stages in the field, a growing number of laboratories that are working on them, and an enhanced appreciation of the role these early life stages play in kelp forest ecology. While this has helped unlock the ‘black box’ surrounding what we know about kelp microscopic life stages, it remains unclear how they will be affected by climate change, and what this will mean for the biogeography and resiliency of kelp forests in the future. In this review, I discuss consequential discoveries surrounding these cryptic life stages and how they influence the biogeography of sporophyte populations that make up these iconic forests. I begin with a discussion of how kelps disperse over both long (100’s to 1,000’s of kilometers) and short (meters to 10’s to meters) distances, and how this relates to their recovery following deforestation. I follow this with an examination of the potential for these microscopic life stages to delay their development and form a ‘bank of microscopic forms’ that allows them to survive during periods of unfavorable conditions. I then explore how variability in irradiance (PAR, UVA + UVB), temperature, nutrients, salinity, seawater pH, sedimentation, and grazing affects the survival, development, and reproduction of these microscopic life stages, and how this can determine biogeographic patterns of macroscopic sporophytes at both local and regional scales. I end with a discussion of the potential of using microscopic life stages in the conservation of kelp forests. For each of these topics, I discuss how these patterns may change under future ocean conditions as predicted by climate change.

Kelp Forests in a Changing Climate

Anthropogenic activities have raised atmospheric carbon dioxide concentrations from a relatively stable 182–300 ppm over the past 800,000 years (Lüthi et al., 2008) to modern levels exceeding 400 ppm (Mauna Loa Observatory, NOAA). Future projections estimate that atmospheric CO2 will surpass 1000 ppm by the year 2100 (IPCC, 2013). Ocean uptake of this excess CO2 (Sabine et al., 2004) and the resulting atmospheric heat is causing the oceans to become more acidic (Doney et al., 2009), warmer (Levitus et al., 2005; Scheffer et al., 2006), and less productive (Gregg et al., 2005; Polovina et al., 2008). Further, as atmospheric and ocean temperatures continue to rise, glacial ablation and enhanced precipitation are increasing freshwater discharge into coastal environments, which is lowering seawater salinities (Boyer et al., 2005; Dyurgerov and Meier, 2005; Arendt et al., 2009; Bieniek et al., 2014; Reisdorph and Mathis, 2014), altering circulation patterns and biogeochemical fluxes (Neal et al., 2010; O’Neel et al., 2015), and intensifying the effects of ocean acidification (Reisdorph and Mathis, 2014). These impacts are especially strong in mid to high latitudes (Miller et al., 2010; Doney et al., 2012) where the nearshore ecosystems are dominated by large forests of kelp (brown algae in the Order Laminariales) that provide food and habitat for numerous other organisms (Konar et al., 2015; Teagle et al., 2017; Metzger et al., 2019; Gabara et al., 2021), enhance primary production (Miller et al., 2011; Edwards et al., 2020; Spector and Edwards, 2020; Sullaway and Edwards, 2020), uptake and store inorganic carbon (Wilmers et al., 2012), alter seawater chemistry (Gonzales et al., 2017; Pfister et al., 2019; Corrano et al., 2020; Corrano et al., 2021), regulate nutrient fluxes (Jackson, 1977), and modulate hydrodynamic activity (Jackson and Winant, 1983; Hondolero and Edwards, 2017). As with other seaweeds (e.g. Ulva; Bews et al., 2021), kelps sequester heavy metals (Evans and Edwards, 2011) and nutrients (Kim et al., 2015) into their tissues, and can therefore be used for bioremediation of polluted waterways. They are also harvested for their nutritional and industrial properties (Borras-Chavez et al., 2012; Borras-Chavez et al., 2016). Consequently, a more detailed understanding how these forests are affected by climate change stressors will be fundamental to predicting patterns of their distribution and abundance, and how coastal ecosystems will function and be used in the future.

Kelp forests have been experiencing broad scale declines in distribution and abundance, and shifts in their latitudinal ranges in many areas of the world due to climate change (Connell and Russell, 2010; Wernberg et al., 2010; Filbee-Dexter et al., 2016; Krumhansl et al., 2016; Assis et al., 2017; Provost et al., 2017; Assis et al., 2018; Beas et al., 2020; Smale, 2020; Wernberg et al., 2019a). They are also subject to deforestation on local scales from a variety of factors, including large storm-driven waves (Ebeling et al., 1985; Seymour et al., 1989; Cavanaugh et al., 2011), El Niño Southern Oscillations (ENSOs) (Tegner and Dayton, 1987; Edwards, 2004; Edwards and Estes, 2006; Edwards, 2019), marine heat waves (Reed et al., 2016; Wernberg et al., 2016; Cavanaugh et al., 2019; Rodgers-Bennet and Catton, 2019; McPherson et al., 2021), sewage discharges and/or spills (Tegner et al., 1995; Connell et al., 2008; Foster and Schiel, 2010), and overgrazing by herbivorous urchins (Estes et al., 1998; Scheibling et al., 1999; Jeon et al., 2015). Indeed, modeling exercises suggest that populations of the kelp Macrocystis pyrifera have a 60% chance of being completely lost from many locations along the California coast, USA during any 20-year period (reviewed in Reed et al., 2006). Given several of these environmental stressors can occur simultaneously, it can be difficult to identify a single stressor responsible for driving kelp loss (Edwards, 2004; Wernberg et al., 2010). Evaluating how these forest recover from these losses can be even more difficult, but doing so will be necessary to predict how kelp forests will ultimately fare under a changing climate.

Kelps exhibit heteromorphic life histories that alternate between microscopic and macroscopic life stages (North, 1994). Specifically, they follow a diplo-haplontic life history in which large diploid sporophytes release microscopic haploid zoospores from reproductive sori on either vegetative or specialized reproductive blades (i.e. sporophylls). These zoospores then disperse across a range of distances and settle on the benthos where they undergo gametogenesis and become either male or female haploid microscopic gametophytes. These gametophytes then produce sperm and eggs, which undergo syngamy to produce embryonic diploid sporophytes that grow in to the next generation of large sporophytes. Generally, the gametophytes need to be in close proximity (i.e. 1mm or closer) to each other to allow the sperm to swim to the eggs for sexual reproduction to occur (Reed et al., 1997). To aid this, female gametophytes release pheromones such as lamoxirene and desmarestene, which can be the main components in egg secretions and cause the male gametophytes to release sperm (Maier et al., 2001). The sperm then exhibit chemotaxis, allowing them to swim to and fertilize the eggs. Although lamoxirene has been identified as the most common pheromone produced in several European and North Pacific kelp species in the genera Laminaria, Alaria, Undaria and Macrocystis, some species such as Laminaria digitata may rely on desmarestene as a more potent chemoattractant (Maier et al., 2001). This suggests that while pheromone-induced sperm release and attraction is conserved among different kelp species, species-specific diversification of complex egg secretions and pheromone receptors can be observed at the chemoattraction level (Maier et al., 2001).

While most studies of kelp forest disturbance and recovery have focused on the large diploid sporophyte stage (e.g. Dayton et al., 1984; Ebeling et al., 1985; Seymour et al., 1989; Edwards, 2004; Edwards and Estes, 2006; Cavanaugh et al., 2011; Cavanaugh et al., 2019; Edwards, 2019), numerous studies have shown convincingly that the haploid microscopic life stages play a critical role in population recovery and in establishing spatial patterns in the larger sporophytes (e.g. Pierce and Cowling, 1991; Ladah et al., 1999; Swanson and Druehl, 2000; Ladah and Zertuche-González, 2007; Wernberg et al., 2019b; Beckley and Edwards, 2021). Given that these microscopic life stages could serve as population bottlenecks (sensu Underweood and Fairweather, 1989), it is important to understand how they respond to environmental variability. Unfortunately, their microscopic sizes and cryptic nature has made them difficult to study in the field (Hsiao and Druehl, 1973; Dayton, 1985; Edwards, 1999; Wernberg et al., 2019b), leading to a black box surrounding what we know about their ecologies and their roles in kelp forest resiliency. However, following Neushul’s (1972) development of a subtidal microscope, advances in genetic analyses of the macroscopic sporophytes (e.g. Coyer et al., 1997; Swanson and Druehl, 2000; Macaya and Zuccarello, 2010; Carney et al., 2013), eDNA metabarcoding of biofilms on natural and biogenic substrates (e.g. Fox and Swanson, 2007; Rubechon et al., 2014; Peters et al., 2015; Akita et al., 2019; Akita et al., 2020a; Akita et al., 2020b), and in experimental field-based techniques that include the use of fluorescent markers (e.g. Cole, 1964; Hsiao and Druehl, 1973; Edwards, 1999; Edwards, 2000), spectrophotometric evaluation of their photosynthetic pigments (e.g. Graham, 1999; Graham and Mitchell, 1999), the collection and laboratory culturing of natural (e.g. Silva, 1992) and outplanted (e.g. Hoffman and Santelices, 1991) substrates, and the sterilization of reef patches to remove the microscopic forms (e.g. Edwards, 1999; Edwards, 2000; Carney et al., 2013) have allowed for a more detailed study of them in the field. Perhaps more informative, however, is their examination in the laboratory under controlled conditions, which has provided an in-depth evaluation of their responses to environmental variability (reviewed in Veenhof et al., in press; Carney and Edwards, 2006; Schiel and Foster, 2015, this paper). As a consequence, the number of published studies on environmental effects on kelp microscopic stages has grown considerably over the past two decades (reviewed in Veenhof et al., in press), which has helped unlock the black box surrounding what we know about them.

Zoospore Dispersal

A fundamental hurdle to understanding how kelp forests colonize newly available habitats is to discern how they disperse over long distances (Hoffman, 1987; Reed et al., 1988; Reed et al., 1992; Reed et al., 2000; Reed and Schroeter, 2004; Batista et al., 2018). For instance, some kelps that possess gas-filled portions of their thalli (e.g. in the genera Macrocystis, Nereocystis, Eualaria, Egregia, and Pelagophycus) may be able to disperse over distances of 100’s to 1000’s of kilometers or more when reproductive sporophytes are dislodged from the substrate and become floating rafts that are transported long distances via offshore currents (reviewed in Thiel, 2003). Such long-distance dispersal from detached macroalgae is believed to have allowed some algae to colonize oceanic islands (reviewed in Van den Hoek, 1987), and may be important in the face of climate change if it increases connectivity among disparate locations, or it opens new areas for colonization (Molinos et al., 2017; Batista et al., 2018). Indeed, floating rafts of Macrocystis pyrifera are commonly observed along the west coasts of North and South America (Dayton et al., 1984; Dayton, 1985; Hobday, 2000; Macaya et al., 2005; Hernández-Carmona et al., 2006; Rothäusler et al., 2009; Hinojosa et al., 2010; Macaya and Zuccarello, 2010), where they have been identified as important long distance dispersal vectors (Batista et al., 2018). Similarly, drifting Nereocystis leutkeana with reproductive sporophylls have been observed cast on the beach and floating near the coast of Shemya Island in the Aleutian Archipelago, which is approximately 1,170 kilometers to the west of the western range limit of the species at Unmak Island, Alaska (Miller and Estes, 1989). However, Nereocystis leutkeana has yet to establish large sporophytes to the west of Unmak Island, which may be due to constraints posed on its microscopic life stages (Miller and Estes, 1989; discussed below). Nereocystis leutkeana sporophytes have also been observed drifting along the coast of Oregon, USA where they have been identified as an important vector for the colonization of associated flora and fauna (Kidder, 2006). In contrast, numerous kelp species do not possess gas-filled thalli, but at least some species (e.g. Ecklonia radiata) may exhibit long-distance dispersal via drifting thalli if they co-occur with other buoyant species such as Sargassum spp. (reviewed in Wernberg et al., 2019b). Still, others may possess gas-fill stipes in parts of their geographic ranges but not in others, as observed in Eisenia arborea (Matson and Edwards, 2006), though it is unknown if this allows for floatation after dislodgement.

Long distance dispersal from floating thalli may be important to the recovery of kelp forests following widespread deforestation. For example, Macrocystis pyrifera populations exhibited widespread losses over several hundred kilometers along Baja California, MEX during the 1997-98 ENSO (Ladah et al., 1999; Edwards, 2004; Edwards and Hernández-Carmona, 2005; Edwards and Estes, 2006; Ladah and Zertuche-González, 2007; Edwards, 2019). These forests recovered rapidly at some locations but took several years to recover at others (Edwards and Estes, 2006). However, it is unclear if recovery of these populations was facilitated by long distance dispersal of floating thalli. It is also unclear if long distance dispersal from floating thalli will be important to the recovery of the Nereocystis leutkeana populations that have been lost over large areas of the coast of northern California, USA following a marine heatwave and a strong ENSO (Rodgers-Bennet and Catton, 2019) or the Eualaria fistulosa populations that have been decimated by urchin grazing throughout most of the Aleutian Archipelago, USA (Estes et al., 1998; Konar et al., 2014; Metzger et al., 2019). Similar patterns of local and widespread kelp loss have been observed in Western Australia (Vanderklift and Wernberg, 2008), South Korea (Jeon et al., 2015), Maine, USA (Steneck et al., 2002), Nova Scotia, CAN (Scheibling et al., 1999), Spain (Voerman et al., 2013), Norway (Fagerli et al., 2013), British Columbia, CAN (Spindel et al., 2021), and central and southern California, USA (Ebeling et al., 1985; Pearse and Hines, 1979; Parnell, 2015) (also reviewed in Krumhansl et al., 2016). Long distance dispersal from floating thalli can be especially important in the face of climate change if changes in wind patterns, ocean heat balances, and/or freshwater inputs from glacial melting and river discharges alter ocean current patterns (e.g. Neal et al., 2010; Sun et al., 2012; Cetina-Heredia et al., 2015; O’Neel et al., 2015; Hays, 2017; Voosen, 2020) and thus influence kelp raft routs and/or speeds. Therefore, a better understanding of the role of long-distance dispersal from reproductive drifting thalli, which remains equivocal (discussed in Reed et al., 1992), and whether this will be altered by climate change is important if we are to evaluate how kelp populations will recover from these losses and persist in the future. Thus, more work in this area is sorely needed.

On a local (i.e. reef) scale, kelp dispersal is predominantly carried out by microscopic zoospores that are passively transported via currents or wave orbitals (Dayton et al., 1984; Reed et al., 1988; Graham, 2003). While most studies suggest that the bulk of successful zoospore dispersal is generally limited to within a few meters of the parental sporophytes (Anderson and North, 1966; Reed et al., 1988; Reed et al., 1992; Reed and Schroeter, 2004), reports based on modelling exercises (Gaylord et al., 2002; Gaylord et al., 2006), genetic analyses of the sporophytes they produce (Coyer et al., 1997; Carney et al., 2013; reviewed in Wernberg et al., 2019b), and the examination of glass microscope slides that had been placed at increasing distances from spore sources (e.g. Reed et al., 1988) have concluded that individual kelp zoospores can effectively disperse over distances of hundreds of meters to several kilometers. In the most extreme case, Amsler and Searles (1980) suggest that zoospores of Macrocystis pyrifera can disperse up to 35 km. Such long distance dispersal of individual zoospores can be facilitated through a variety of mechanisms, including the greater production and/or synchronous release of zoospores from specific locations within or near the edge of a forest (Reed et al., 1988; Amsler and Neushul, 1989a; Norton, 1992; Reed et al., 1997; Graham, 2003; Edwards and Konar, 2012), the vertical transport of zoospores into shallower portions of the water column where current velocities and dispersal potentials are greater (Amsler and Searles, 1980; Stevens et al., 2003; Cie and Edwards, 2011), and through processes that slow their sinking speeds (Hoffman and Camus, 1989; Gaylord et al., 2002; Raimondi et al., 2004). The latter may be aided by the fact that kelp zoospores possess large lipid reserves and are able to photosynthesize and swim while in the water column (Reed et al., 1992; Brzezinski et al., 1993). For example, the zoospores of Macrocystis pyrifera can swim for more than 120 hrs, though the majority of them likely stop swimming after about 48 hrs (Amsler, 1988; Amsler and Neushul, 1991; Reed et al., 1997). In contrast, the zoospores of Ecklonia radiata are able to swim for at least 24 hours but most appear to swim for only 1 or 2 hours (reviewed in Wernberg et al., 2019b), and few zoospores of Laminaria hyperborea likely swim for longer than 20 hours (Kain, 1964). However, swimming may affect their ability to settle, though this appears to vary among species. For example, Macrocystis pyrifera zoospores lose their ability to settle soon after they stop swimming, while Pterygophora californica zoospores appear to increase their ability to settle after they stop swimming (Reed et al., 1992). If the zoospores do not settle on the benthos, they can still form gametophytes, as observed in Lessonia nigrescens whose zoospores have been observed germinating after 4 days in the water column (Hoffman and Camus, 1989). It remains unclear, however, if these gametophytes can then settle on the substrate.

Even with the ability to swim for extended periods, kelp zoospores are likely too small (~3-7 μm) and their swimming speeds too slow (~0.0012 mm s-1) to effectively disperse (Gaylord et al., 2002). Rather, swimming may simply allow the zoospores to find suitable substrates once in the boundary layer using chemotaxis (Amsler and Neushul, 1989b), or it may help move the zoospores vertically out of the boundary layer to prevent settlement, which can increase their time in the water column and thereby increase their current-driven dispersal potential (Amsler and Searles, 1980; Reed et al., 1988; Amsler et al., 1992; Reed et al., 1992). However, as zoospore dispersal distances increase, their density in the water column decreases (Gaylord et al., 2002), resulting in diminishing settlement densities (Reed et al., 1997). This is important given the need for zoospores to be in close proximity so that the resulting male and female gametophytes are close to each other for successful sexual reproduction to occur (Dayton, 1985; Reed et al., 1997). On the other hand, increasing dispersal distances can also effectively reduce the density of related zoospores (i.e. siblings) settling next to each other, and thus decrease the negative effects associated with inbreeding (Raimondi et al., 2004; Carney et al., 2013). However, the negative effects of limited dispersal may be lessened if these species have already purged deleterious recessive alleles from their haploid stages, thereby allowing selfing to be an effective reproductive strategy as observed in Postelsia palmaeformis (Barner et al., 2010). Selfing may also allow species to more effectively colonize new habitats following long-distance dispersal given only a single individual needs to successfully establish there (Baker, 1955).

The transport of zoospores into shallower water can result in their being exposed to greater irradiances and higher planktonic grazer abundances, both of which reduce their survival and settlement competency (Cie and Edwards, 2008; Müller et al., 2009; VanMeter and Edwards, 2013). For example, VanMeter and Edwards (2013) exposed the swimming zoospores of Macrocystis pyrifera to mysids, which are small shrimp-like planktonic grazers that are abundant in southern California, USA kelp forests (Coyer, 1984; Turpen et al., 1994). They then used fluorescence microscopy to identify chlorophyll in the mysid guts and verified that they did in fact consume the zoospores, which in turn resulted in significant reductions in zoospore settlement relative to conditions without mysids. This is important given these mysids are more abundant in the shallower portions of the kelp canopies than near the benthos (Coyer, 1984; VanMeter and Edwards, 2013). This has obvious implications for the future if climate change alters the abundance of planktonic grazers in these forests, which may be expected given that changes in the distribution and phenology of zooplankton are occurring in numerous places in the world (e.g. Richardson, 2008; Johnson et al., 2011; Dam and Baumann, 2017). High irradiance (PAR, UVA+UVB) also negatively affects kelp zoospores, but these effects vary among kelp species (Swanson and Druehl, 2000; Wiencke et al., 2000; Roleda et al., 2005; Cie and Edwards, 2008; reviewed in Bischof et al., 2006). For example, Cie and Edwards (2008) exposed the swimming zoospores of Macrocystis pyrifera and Pterygophora californica to a range of irradiances between 75 and 1050 μmol photons m-2s-1 PAR for periods of time between 1 and 12 hrs and found that while settlement of Macrocystis pyrifera was not significantly affected by exposure to high irradiance, settlement of Pterygophora californica was significantly reduced, and this effect strengthened with longer exposure times. Indeed, Pterygophora californica settlement completely ceased when its zoospores were exposed to irradiances of ≥575 μmol photons m-2s-1 PAR for 12 hrs. Further, exposure to high irradiances resulted in delayed effects to the benthic microscopic life stages following settlement, as Macrocystis pyrifera zoospores did not produce viable gametophytes or embryonic sporophytes following settlement when they were exposed to irradiance of 75 μmol photons m-2s-1 PAR for 12 hrs, or to irradiances of 1025 μmol photons m-2s-1 PAR for as little as 4 hrs. In contrast, Pterygophora californica zoospores produced gametophytes and embryonic sporophytes under all irradiances and exposure times tested, although the density of these decreased with increasing irradiance and exposure times. Interestingly, the negative effects of high irradiance often do not appear in the adult sporophytes (Edwards and Kim, 2010; Fejtek et al., 2011). Rather, this appears to be related to the light harvesting characters associated with photosystem II in the zoospores, which Graham (1999) found to be significantly different among even closely related kelp species. This difference is sufficient to discriminate among species of kelp zoospores using micro spectrophotometry (Graham, 1999; Graham and Mitchell, 1999). Regardless of their tolerances to high irradiances, shading from the kelp canopies can decrease irradiance within kelp forests and thus reduce these negative effects, as seen for Alaria esculanta along the coast of Kongsforden, Svalbard in the Arctic Ocean (Laeseke et al., 2019). Further, some kelps (e.g. Nereocystis leutkeana) may avoid exposing their zoospores to high irradiances by releasing them within a few hours just before or after sunrise (Amsler and Neushul, 1989a).

The negative effects of high irradiance on zoospores may be important in establishing patterns of depth distributions in the large sporophyte populations. For example, Swanson and Druehl (2000) measured zoospore tolerance to UV light in Macrocystis pyrifera, Pterygophora californica, Saccharina groenlandica, and Hedophyllum sessile within Barkley Sound, British Columbia and found that it correlated with sporophyte depth zonation patterns for each species. Likewise, Wiencke et al. (2000) found that when the zoospores from different species of kelp in Spain and Norway were exposed to similar levels of UV light, the zoospores from species occurring in deeper water exhibited lower germination rates than those obtained from species in shallower water. These negative effects, however, varied depending on the quality (wavelength) of light the zoospores experienced. Specifically, zoospores exhibited higher mortality when exposed to PAR+UVA+UVB than when exposed to PAR+UVA or PAR alone. Wiencke et al. (2000) also concluded that the loss of zoospore viability in species such as Laminaria digitata was primarily due to photo damage to the zoospores’ DNA and photosystems. Similar findings from Helgoland, North Sea, have shown that zoospores obtained from deeper-water kelps such as Laminaria hyperborea are more strongly affected by UV light than the zoospores from shallower kelps such as Laminaria digitata and Saccharina latissima (Roleda et al., 2005), and that recovery of damaged photosystems was also related to their depth distributions, as Laminaria digitata exhibited greater recovery than either Saccharina latissima or Laminaria hyperborea. However, it should be noted that many kelp genera (e.g. Nereocystis, Ecklonia, Eisenia) do not fit obvious depth gradients, but rather span much of the water column and therefore may not be as reliant on zoospore tolerances to high irradiance. Regardless, the effects of irradiance can have important implications for the future if climate change results in reduced productivity in coastal waters due to “coastal darkening” (Blain et al., 2021; Wollschläger et al., 2021) and/or greater losses of the kelp canopies due to increased storm activity (e.g. Seymour et al., 1989; Justic et al., 1997; Scavia et al., 2002; Behrenfeld et al., 2006; Byrnes et al., 2011; Bakun et al., 2015). Both of these can result in altered subtidal irradiances (Edwards, 1998; Clark et al., 2004; Foden et al., 2010) and thus altered patterns of zoospore production (Reed et al., 1988; Edwards and Konar, 2012). These impacts may be further exacerbated if increases in ocean acidification (OA) and/or ocean warming (OW) from climate change also negatively affect kelp zoospores. For example, Hoos (2015) showed that exposure to OA and OW resulted in mechanical and functional damage to the zoospores of Egregia menziesii that ultimately decreased their swimming speeds, settlement densities, adhesion abilities, and germination rates. Hoos (2015) concluded that these negative effects combined with reductions in habitat space for the macroscopic sporophytes will ultimately result in reductions in the abundance of future Egregia populations. Likewise, Shukla and Edwards (2017) found that exposing the zoospores of Macrocystis pyrifera to the individual effects of OA and OW resulted in decreased zoospore settlement and subsequent lower gametophyte production. Specifically, gametophyte production was four times greater when the swimming zoospores were exposed to 12°C than 15°C, and 25% greater when the zoospores were exposed to 400 µatm pCO2 seawater than when they were exposed to 1500 µatm pCO2 seawater. Given OA and OW are expected to increase in the future, this can have significant consequences on patterns of kelp abundance and distribution.

Do Banks of Microscopic Life Stages Form a ‘Seed Bank Analogue’?

Once kelp zoospores have dispersed, they can settle on the benthos and germinate into male and female gametophytes. Settlement location appears to be non-random and selected for by the zoospores. For example, Macrocystis pyrifera and Pterygophora californica settlement is stimulated by nutrients, which may allow them to either choose or avoid microhabitats that favor or inhibit growth and reproduction in the resulting gametophytes (Amsler and Neushul, 1990). Further, selection of microhabitats may be aided by the zoospores actively swimming towards or away from microhabitats using chemotaxis (Amsler and Neushul, 1989b), and they can actively select surface depressions that facilitate survival once they are within the boundary layer (Amsler et al., 1992). Such surface depressions can then aggregate zoospore settlement, which will reduce the distance between gametophytes and enhance reproduction and sporophyte recruitment (Muth, 2012). Following settlement and germination, the resulting gametophytes can either undergo sexual reproduction and produce sporophytes, or they can delay reproduction and persist as a ‘bank of microscopic forms’ if environmental conditions are unfavorable to sporophyte growth and survival (Klinger, 1984; Chapman, 1986; Hoffman and Santelices, 1991; Blanchette, 1996; Ladah et al., 1999; Carney and Edwards, 2006; Ladah and Zertuche-González, 2007; Carney and Edwards, 2010; Carney, 2011). More broadly, organisms that live in environments where conditions are temporally variable can reduce their metabolisms (Crowe, 1971; Pinter et al., 1984; Geiser, 2004; Heldmaier et al., 2004; Guidetti et al., 2011; Tøien et al., 2011; Careau et al., 2014; Dolinar and Edwards, 2021), or they can rely on alternate life stages that are either more tolerant of, or remain dormant during, periods of unfavorable conditions (Tauber and Tauber, 1978; Lubchenco and Cubit, 1980; Slocum, 1980; Hochachka and Guppy, 1987). These organisms can then emerge from dormancy and resume metabolic activity when conditions again become favorable for growth and survival (Hinton, 1968; Hollibaugh et al., 1981; Dolinar and Edwards, 2021). Indeed, the reliance on dormant life stages has been observed across a wide range of organisms, including terrestrial plants (Venable and Lawlor, 1980; Keeley, 1987; Leck et al., 1989), insects (Tauber and Tauber, 1978), microalgae (Hollibaugh et al., 1981), and marine crustaceans (Grice and Marcus, 1981; Maier, 1990). They have also been proposed for kelps and other marine macroalgae with similar diplo-haplontic life histories (Dayton, 1973; Dayton, 1985; Chapman, 1986; Hoffman and Santelices, 1991; Edwards, 2000; Kinlan et al., 2003; Carney and Edwards, 2006; Carney and Edwards, 2010; Carney, 2011; Carney et al., 2013; Ebbing et al., 2020; Schoenrock et al., 2021) where they have been shown to be important to sporophyte recruitment, especially when no sources of new zoospores are available (Dayton, 1973; Silva, 1992; Carney et al., 2005). For example, Silva (1992) collected biogenic substrates from deep water off Cordell Bank, California, USA and cultured them in the laboratory. He observed the recruitment of a Nereocystis leutkeana sporophyte on this substrate even though no apparent source of zoospores could be identified, and concluded that the sporophyte emerged from a vegetative gametophyte that had survived for several years. Likewise, Edwards (2000) sterilized the substrate in experimental plots on a rocky reef in Carmel Bay, California to remove all seaweed microscopic stages after all Desmariesia ligulata (a brown alga with a similar diplo-haplontic life history) sporophytes had disappeared from the reef due to natural senescence and winter storms, and found that this prevented new sporophyte recruitment in the following spring. In contrast, a dense recruitment of Desmarestia sporophytes was observed in adjacent non-sterilized plots. Lastly, Ladah and Zertuche-González (2007) suggest that microscopic life stages of Macrocystis pyrifera were able to survive the 1997-98 ENSO near their southern range limit in Baja California, MEX, perhaps in deeper water, and then promote kelp forest recovery following the ENSO.

Although numerous studies have suggested the importance of delayed development in kelp microscopic life stages, it remains unclear which life stage (haploid gametophytes, diploid sporophytes, or both) undergoes this delay. While the haploid gametophytes have most often been suggested as the life stage that undergoes delayed development (Kain, 1964; Dayton, 1973; Hsiao and Druehl, 1973; Klinger, 1984; Dayton, 1985; Silva, 1992; Blanchette, 1996; Edwards, 1999; Ladah et al., 1999; Edwards, 2000; Carney and Edwards, 2010; Ebbing et al., 2020), some studies have suggested that diploid embryonic sporophytes may serve as the stage that delays development (Kinlan et al., 2003; Ladah and Zertuche-González, 2007). This distinction is important given that both male and female gametophytes need to survive in close proximity to each other to undergo sexual reproduction once they resume development. However, delaying development as diploid embryonic sporophytes can allow for more rapid sporophyte recruitment when conditions become favorable because only one individual (the sporophyte) needs to survive the delay period, and because sporophyte recruitment is not slowed by the need to undergo sexual reproduction after their period of delay, as reproduction has already occurred. This, therefore may offer a competitive advantage by producing macroscopic sporophytes more quickly than in any competitors (Carney and Edwards, 2010; Carney, 2011). Diploid stages may also be better adapted to a broader range of environmental conditions than their haploid counterparts due to a greater array of possible gene combinations, which may serve as protection against the expression of deleterious mutations, and thus are often selected for in macroalgae (Perrot et al., 1991; reviewed in Thornber and Gaines, 2004). For example, Muth et al. (2021) found that the microscopic diploid sporophytes of Lamniaria solidungula in the Beaufort Sea were more tolerant to decreased salinities than the haploid gametophytes, while Matson and Edwards (2007) similarly showed that the large sporophytes of Eisenia arborea along the west coast of California, USA and Baja California, MEX were more tolerant to elevated temperatures than the microscopic haploid gametophytes. In contrast, haploid gametophytes have lower mutational loads (Crow and Kimura, 1965) and lower nutritional requirements because they have half the DNA content (Lewis, 1985; reviewed in Thornber and Gaines, 2004). Although studies identifying the life stage that makes up the bank of microscopic stages are few, the outplanting of glass slides containing the microscopic stages of Nereocystis leutkeana (Hsiao and Druehl, 1973) and Desmarestia ligulata (Edwards, 2000) that had been stained with calcofluor white and then later collected and examined using fluorescence microscopy has identified microscopic gametophytes as the life stage that forms the banks of microscopic stages in the field.

The process of delaying development in kelp gametophytes can be triggered when they do not receive adequate nutrients (Carney and Edwards, 2010), levels of iron (Lewis et al., 2013), or light (i.e. photosynthetically usable radiation; PUR) (Ebbing et al., 2020). In general, blue light has been shown to be important to proper algal development (e.g. Dring, 1988). More specifically, depriving kelp gametophytes of blue light can result in their failing to undergo sexual reproduction, and instead may allow them to persist for long periods of time. Indeed, haploid gametophytes can survive in a vegetative state in the laboratory for years and still produce sporophytes when returned to conditions that favor sexual reproduction (Yoneshigue-Valentin, 1990; Hoffman and Santelices, 1991; tom Dieck, 1993; see also reviews in tom Dieck, 1993; Carney and Edwards, 2006; Barrento et al., 2016). Delaying reproduction, in turn, may be beneficial if it allows for more rapid sporophyte recruitment when favorable conditions return. For example, Carney (2011) found that delayed development in the microscopic gametophytes of Macrocystis pyrifera, Pterygophora californica, Pelagophycus porra and Laminaria farlowii resulted in more rapid sporophyte production compared with recently settled zoospores, perhaps conferring a competitive advantage for these sporophytes. Haploid gametophytes have consequently been characterized as ‘seed bank analogues’ (Edwards, 2000; Ladah and Zertuche-González, 2007) that are of mixed ages and different parentages, which can reduce inbreeding (Carney et al., 2013). However, unlike plant seeds, kelp gametophytes are haploid and consist of dioecious males and females that can exhibit different reproductive strategies. For example, Destombe and Oppliger (2010) note that the male gametophytes of Laminaria digitata grow and reproduce simultaneously, while females stop growing after reproduction, which they suggest enhances the reproductive success of the species. However, it is important to note that some studies that have placed boulders within kelp forests to collect zoospores for different periods of time suggest that kelp microscopic stages have little capacity for long-term survival in nature, at least in Macrocystis pyrifera and Pterygophora californica, and that sporophyte recruitment is largely driven by newly settled zoospores (Reed et al., 1997). This is similar to observations for some terrestrial plant seed banks when sources of new seeds are available (e.g. Alvarez-Buylla and Martínez-Ramos, 1990; reviewed in Cohen and Levin, 1987).

Unlike the seeds of terrestrial plants that remain truly dormant during stressful periods, the microscopic life stages of marine algae are physiologically active and highly sensitive to even small changes in environmental conditions (Graham, 1996; Matson and Edwards, 2007; Fredersdorf et al., 2009; Fejtek et al., 2011; Martins et al., 2017), especially regarding the interactive effects of climate change (Gaitán-Espitia et al., 2014; Shukla and Edwards, 2017). In particular, they are known to be sensitive to increases in temperature (Lüning, 1980; Ladah and Zertuche-González, 2007; Matson and Edwards, 2007; Martins et al., 2017; Lind and Konar, 2017; Shukla and Edwards, 2017; Small and Edwards, 2021), irradiance and UV (Graham, 1996; Edwards, 2000; Kinlan et al., 2003; Müller et al., 2008; Fredersdorf et al., 2009; Roleda, 2009; Fejtek et al., 2011; Ebbing et al., 2020; Silva et al., 2022), sedimentation (Devinny and Volse, 1978; Carney et al., 2005; Deiman et al., 2012; Zacher et al., 2016; Traiger and Konar, 2017), and grazing (Dean et al., 1984; Leonard, 1994; Henríquez et al., 2011; Zacher et al., 2016), and to decreases in salinity (Fredersdorf et al., 2009; Lind and Konar, 2017; Muth et al., 2021), nutrient availability (Hoffman et al., 1984; Hernández-Carmona et al., 2001; Kinlan et al., 2003; Muñoz et al., 2004; Carney and Edwards, 2010), and ocean pH (Olischläger et al., 2012; Shukla and Edwards, 2017). Temperature, in particular, has been shown to have both positive and negative effects on kelp gametophytes, and these effects can vary among different kelp life history states, as these microscopic stages of at least some kelp species (e.g. Hedophyllum nigripes and Laminaria digitata) appear more tolerant to elevated temperatures that their macroscopic sporophytes are (Franke et al., 2021), but they are less tolerant in other species (e.g. Pterygophora californica) (Matson and Edwards, 2007). Further, the effect of temperature can be species-specific, which will likely affect the distribution and abundance of these forests in the future (e.g. Muth et al., 2019), and perhaps favor one species over another (e.g. Franke et al., 2021). For example, studies on Macrocystis pyrifera (Ladah and Zertuche-González, 2007; Shukla and Edwards, 2017), Nereocystis leutkeana, Eualaria fistulosa, Saccharina latissima (Lind and Konar, 2017), Pterygophora californica (Matson and Edwards, 2007) and Laminaria spp. (Lüning, 1980) have shown elevated temperatures that result from ocean warming or that occur along latitudinal gradients negatively affect growth and development in their gametophytes (Shukla and Edwards, 2017), while other studies on Alaria esculenta and Laminaria digitata (Silva et al., 1992) have shown elevated temperatures that arise due to seasonality positively affecting these stages. For the latter, it seems that cold temperatures in the winter in high latitude ecosystems may inhibit gametophyte development, but when temperatures warm to their summer levels, the gametophytes undergo sexual reproduction and produce sporophytes (Silva et al., 1992). This may have complex life history implications, as Martins et al. (2017) revealed that elevated temperatures cause Laminaria digitata gametophytes to delay their reproduction and thus grow larger, which may result in enhanced sporophyte recruitment in the autumn and spring. Further, elevated temperatures may cause complex changes in the way kelp gametophytes are affected by other abotic and biotic stressors such as sedimentation and grazing, which can lead to shifts in community structure in the future, as suggested for Alaria esculanta, Laminaria digitata, and Saccharina latissima in the Arctic (Zacher et al., 2016).

Kelp microscopic stages are sensitive to inter and intra specific competition due to differential settlement densities (Reed, 1990; Carney and Edwards, 2010), which can be modified by changes in ocean temperatures. For example, Zacher et al. (2019) found that ocean warming positively affected spore germination, gamtogenesis and sporophyte formation in Alaria esculenta and Laminaria digitata in the Arctic, and that warming resulted in enhanced development in Alaria esculenta when co-cultured with Laminaria digitata, but this was not observed under colder (non-warming) conditions. Zacher et al. (2019) concluded that ocean warming could allow Alaria esculenta to gain a competitive advantage over Laminaria digitata and thereby alter community structure in Arctic the future. Given many of the factors associated with climate change are expected to change in the coming decades (Scavia et al., 2002; Boyer et al., 2005; Dyurgerov and Meier, 2005; Levitus et al., 2005; Behrenfeld et al., 2006; Arendt et al., 2009; Doney et al., 2009; Bieniek et al., 2014; Reisdorph and Mathis, 2014), it is likely that they, and more importantly their interactions, will be important to establishing and/or maintaining biogeographical patterns in the large sporophytes that make up the kelp forests. However, because the manner in which kelp microscopic stages respond to these stressors varies among kelp species, climate change may drive shifts in species compositions in some locations (e.g. Deiman et al., 2012; Lind and Konar, 2017; Traiger and Konar, 2017; Silva et al., 2022). Such changes have already been seen for Laminaria ochroleuca and Laminaria hyperborea, with the more warm tolerant species Laminaria ochroleuca increasing in abundance and distribution along the rapidly warming Western English Channel (Smale et al., 2015). Likewise, Hondolero and Edwards (2017) note changes in the relative abundance of two kelp species, Eualaria fistulosa and Nereocystis leutkeana, in Kachemak Bay, Alaska, USA, with the more warm tolerant Nereocystis leutkeana increasing in abundance and Eualaria fistulosa decreasing in abundance. Indeed, some models predict that Northern hemisphere kelps will migrate northwards as these waters become more suitable for their growth and survival in the future (Assis et al., 2018; Smale, 2020). Therefore, a better understanding of how kelp benthic microscopic life stages are affected by environmental stressors and how this affects the way these stages interact with each other is key to better predicting how these ecosystems will be structured in the future.

The emergence of macroscopic sporophytes from the bank of microscopic forms (i.e. sporophyte recruitment) is strongly affected by complex interactions of numerous environmental factors. For example, recruitment of Desmarestia ligulata sporophytes occurs during a very narrow period in the spring when day lengths are increasing (Edwards, 1998). This is further supported by early work on Saccharina japonica by Tseng et al. (1959) and Saccharina latissima by Lüning (1980) who revealed that the light environment, including photoperiod, irradiance, and light color, is deterministic in the process of gametogenesis. Specifically, these gametophytes exhibit diel rhythms with day-night cycles synchronizing egg release and fertilization being largely limited to dark periods, and with egg release being inhibited by blue and UV light (λ = 372, 413, 438, and 481 nm). Response to changes in irradiance and light color may be important in the future if climate change results in increased water turbidity and thus coastal darkening (Wollschläger et al., 2021), which can affect gametogenesis and thereby alter sporophyte recruitment. Additionally, the process of sporophyte recruitment is generally negatively affected by high irradiance (Wernberg et al., 2019b; Ebbing et al., 2020; Paine et al., 2021; Silva et al., 2022) and elevated temperatures (Izquierdo et al., 2002; Morita et al., 2003; Mohring et al., 2014; Wernberg et al., 2019b; Franke et al., 2021; Paine et al., 2021; Silva et al., 2022), and positively influenced by increases in day length (Dring, 1988; Edwards, 2000; Nelson, 2005; Ratcliff et al., 2017) and elevated nutrient availability (Muñoz et al., 2004; Carney and Edwards, 2010; Carney, 2011; Ratcliff et al., 2017). Specifically, elevated seawater temperatures have been shown reduce gametophyte survival and/or sporophyte production in numerous species of kelp, including Saccharina latissima (Bolton and Lüning, 1982), Pterygophora californica (Matson and Edwards, 2007), Macrocystis pyrifera (Shukla and Edwards, 2017), Ecklonia radiata (Wernberg et al., 2019b), Lessonia corrugata (Paine et al., 2021), Laminaria digitata and Hedophyllum nigripes (Franke et al., 2021), Alaria esculanta and Laminaria digitata (Silva et al., 2022), among others (e.g. Muth et al., 2019). However, the upper thermal tolerance of kelp gametophytes varies among species and with latitude (tom Dieck, 1993; Matson and Edwards, 2007; Muth et al., 2019; Zacher et al., 2019), and between male and female individuals (Franke et al., 2021). As noted earlier, these temperature tolerances may also be greater than those of the larger sporophytes, as seen in Laminaria digitata and Hedophyllum nigripes (Franke et al., 2021), or they may be less tolerant than the larger sporophytes, as seen in Pterygophora californica (Matson and Edwards, 2007). Likewise, OA has also been observed to negatively affect gametophyte survival and sporophyte production in some kelps, such as Macrocystis pyrifera (Gaitán-Espitia et al., 2014; Shukla and Edwards, 2017), but that increased CO2 may ameliorate the negative physiological effects on zoospore germination and result in overall positive effects (Roleda et al., 2012). This again varies among species, as OA has also been observed to accelerate oogonium formation, but not affect sporophyte production in Laminaria hyperborea (Olischläger et al., 2012). Further, as with the macroscopic sporophytes (e.g. Brown et al., 2014), the combined effects of OW and OA appear equivocal, as they have been shown to have both positive (Shukla and Edwards, 2017) and negative (Gaitán-Espitia et al., 2014) effects on growth and development of Macrocystis pyrifera gametophytes, and they may affect the ability of these microscopic stages to delay their development (Gaitán-Espitia et al., 2014). When favorable levels of these conditions occur together, they can create “recruitment windows” during which conditions become ideal to support sporophyte production (Deysher and Dean, 1984; Deysher and Dean, 1986). Indeed, such recruitment windows have been observed in the spring along the west coast of North America when sporophyte recruitment is most abundant (Dayton et al., 1984; Edwards, 1998), but it remains unknown how climate change will affect sporophyte recruitment in the future (Harley et al., 2012). Certainly, this area can benefit greatly from additional studies.

Microbe Associations

To date, studies on multiple canopy-forming kelp species suggest they are associated with distinct microbial communities (Lemay et al., 2018; Lin et al., 2018; Minich et al., 2018; Weigel and Pfister, 2019; Ramirez-Puebla et al., 2020; Phelps et al., 2021). Studies from British Columbia, CAN on Costaria costata, Alaria marginata, Pterygophora californica, Cymathaere triplicata, Laminria setchellii, Nereocystis leutkreana, Saccharina groenlandica, Saccharina latissima, for example, have revealed that their microbiomes are distinct from the water column but are similar across different species of kelp, with annual versus perennial species supporting different microbial communities (Lemay et al., 2018). They have also been shown to vary among different parts of the kelp thalli that are of varied ages (Lemay et al., 2021). Further, kelp-associated microbiomes, which are important to physiological functions (Lemay et al., 2021), can shift across depth gradients (Lin et al., 2018). OA has been shown to alter the microbial communities immediately surrounding some kelps (e.g. Macrocystis pyrifera) (Minich et al., 2018) and nearby benthic coralline algae (e.g. Lithothamnion) (Cavalcanti et al., 2018), and changes in these communities can have significant effects on the survival and development of kelp gametophytes (Morris et al., 2016). However, these effects appear to vary geographically, as Morris et al. (2016) found that the microbial communities found in seawater collected from Point Loma, San Diego, CA had detrimental effects on development in Macrocystis pyrifera gametophytes, while the microbial communities found in seawater collected from Catalina Island, which is 133 km to the northwest, had beneficial effects on the these gametophytes. They attributed this variability to differences in how human populations affect the seawater microbial communities, as Point Loma lies immediately adjacent to the city of San Diego and is heavily impacted by a large human population, while Catalina Island lies approximately 40 km offshore and is more pristine. Therefore, the ultimate effects of climate change on kelp microscopic stages will likely involve complex synergies among multiple stressors, including OW, OA, irradiance, microbial communities, and human influences on the coastal environment. Further, if climate change reduces the frequency or duration of the recruitment windows, it can negatively affect how kelp microscopic stages emerge from their period of delayed development and thereby reduce patterns of sporophyte recruitment and slow recovery following deforestation in the future.

The Importance of Microscopic Stages to Kelp Biogeography

As noted earlier, the individual and combined effects of variability in temperature, irradiance, salinity, large-scale climatic events, and climate change can have profound effects on patterns of kelp biogeography (Buschmann et al., 2004; Graham et al., 2007; Schiel and Foster, 2015; Smale, 2020; Muth et al., 2021). Thermal temperature tolerances, in particular, have a strong influence on establishing patterns among latitudes (Van den Hoek, 1982) and appear to be driving shifts in the geographic distributions of some species (discussed above). However, within latitudes, different kelp species exhibit distinct geographic range limits due to their species-specific requirements for, and/or tolerances of, temperature, nutrient, and light conditions (tom Dieck, 1993; Oppliger et al., 2012; Muth et al., 2019). While some of this occurs through impacts to the macroscopic sporophytes, environmental constraints on the microscopic stages can also be integral in establishing biogeographic patterns in kelp populations (tom Dieck, 1993; Matson and Edwards, 2007; Oppliger et al., 2012; Wernberg et al., 2019b; Muth et al., 2021). Further, while the macroscopic sporophytes of at least some kelp species, such as Macrocystis pyrifera (Kopczak et al., 1991), Saccharina latissima (Gerard, 1988; Gerard, 1990), Eisenia arborea (Roberson and Coyer, 2004), Laminaria digitata (Liesner et al., 2020), and Ecklonia radiata (Bennett et al., 2015) exhibit ecotypic adaptation to local conditions that may allow them to persist where other ecotypes cannot, it is unclear if the microscopic stages do as well (Bolton and Lüning, 1982). However, studies on Saccharina latissima in Long Island, New York, USA (Gerard, 1990) and Ecklonia radiata from New Zealand (Novaczek, 1984) suggest that the microscopic gametophytes of at least some species do exhibit ecotypic adaptation to local conditions. The importance of this, however, largely remains within the black box of uncertainty surrounding kelp microscopic stages.

Regardless of whether kelp microscopic stages exhibit ecotypic adaption to local conditions, there is strong evidence that they can be integral in establishing biogeographic ranges in the forest-forming large sporophytes (tom Dieck, 1993; Muth et al., 2019). For example, Matson and Edwards (2007) note that Pterygophora californica and Eisenia arborea both occur along the west coast of North America and have similar northern ranges near Vancouver Island, British Columbia, CAN. However, their southern ranges differ, with Pterygophora californica’s southern range ending at Bahía Rosario, Baja California, MEX, and Eiseni arborea’s range extending approximately 550 km farther south to Bahía Magdalena, Baja California Sur, MEX. Matson and Edwards (2007) found that while the adult sporophytes of both species are tolerant of the warmer waters observed in the south, the microscopic gametophytes exhibit very different responses to these temperatures. Specifically, in laboratory experiments, Eisenia arborea gametophytes survived and produced embryonic sporophytes under both 12° and 18°C, which are characteristic of bottom temperatures observed in Bahía Rosario and Bahía Magdalena, respectively, but the gametophytes of Pterygophora californica produced embryonic sporophytes only under the cooler temperatures. In fact, the gametophytes of Pterygophora californica all died when grown under 18°C. This is similar to the findings of Small and Edwards (2021) who found that expansion of the southern and northern biogeographic range limits of the invasive brown alga Sargassum hornerni along the California, USA coast were largely set by low tolerance of its microscopic germlings to both high and low temperatures, respectively, as the germlings did not fully develop under temperatures that are outside the species current range. Although Sargassum horneri is not a kelp and does not exhibit a diplo-haplontic life cycle, but rather is in the order Fucales and exhibits a diplontic (animal-like) lifecycle, it provides further experimental evidence that the geographic ranges of habitat-forming seaweeds can be set by abiotic influences on their microscopic life stages.

On a longitudinal gradient, Miller and Estes (1989) observed that the western range limit of Nereocystis leutkeana in the Aleutian Islands was set at Unmak Island, just to the east of the Samalga Pass. While they did not determine the factors that set this range limit, they did hypothesize that it was likely due, at least in part, to the ability of its microscopic stages to reproduce under the lower light conditions that are established by the heavier cloud and fog cover to the west. In suggesting this, they note that there was an abundance of drifting reproductive sporophytes west of the Samalga Pass (discussed earlier). They also note that Nereocystis leutkeana’s microscopic life stages do not develop properly under low light conditions (Vadas, 1972), and that the western Aleutians are generally foggy and characterized by low irradiance in the summer when Nereocystis leutkeana recruitment occurs (Armstrong, 1977). This again may help explain why no Nereocystis leutkeana sporophytes have been observed farther to the west on Shemya Island even though drifting reproductive sporophytes have likely provided spores there. Lastly, Muth et al. (2019) examined the effect of ocean temperature on 12 species of kelp from the eastern Pacific Ocean and found that sporophyte production was always observed at 12°C, but sporophyte failure was common at a warmer temperature of 18°C. Muth et al. (2019) conclude that warming ocean temperatures will likely cause recruitment failure of some species, especially those near the warmer edge of their ranges, and thus affect kelp resiliency in the future. A major exception to this has been observed in two species of Undaria (Undaria pinnatifida and Undaria undarioides), which both exhibited optimal temperatures for gametophyte growth to be 20°C along the coast of Japan (Morita et al., 2003). However, the optimal temperature for gametophyte maturation was 10 – 15°C for Undaria pinnatifida and 20-21°C for Undaria undarioides, and this difference is believed to be a major factor determining the distribution of these two species along the coast of Japan (Morita et al., 2003). In contrast, Henkel et al. (2008) studied gametophytes of Undaria pinnatifida that were obtained from different locations along the California coast, USA and found that they exhibited very broad thermal tolerances, remaining metabolically active in temperatures up to 31°C, which exceeded normal environmental conditions. Further, Watanabe et al. (2014) found that gametophytes of Undaria pinnatifida collected from Kagoshima Bay, Japan could not survive temperatures of 28°C, and that temperatures above 20°C affected gametophyte performance. Watanabe et al. (2014) concluded that the species could become locally extinct along the coast of Japan if temperatures continue to rise with climate change.

The Importance of Kelp Microscopic Stages to Patterns of Depth Zonation

Kelps occur from the intertidal to depths of more than 30 m (Schiel and Foster, 2015), with some reports of their existence to more than 200m (Žuljević et al., 2016). While some of the depth patterns in kelps can be attributed to irradiance effects on their swimming zoospores, much can also be attributed to effects on its benthic microscopic life stages, as discussed above. For example, populations of Macrocystis pyrifera appear to be excluded from the intertidal and very shallow subtidal along much of the California, USA coast due to hydrodynamic forces that act on its large sporophytes (Graham, 1997) and to high irradiances that negatively affect its microscopic stages (Graham, 1996). Specifically, Graham (1996) examined the effect of high irradiance on the benthic microscopic stages of Macrocystis pyrifera and found that the species’ shallow water limit was set, at least in part, by the inability of its microscopic stages to tolerate high irradiance. While he did not identify the life stage that was responsible, this is in agreement with earlier work by Lüning and Neushul (1978) who found that exposing the gametophytes of Macrocystis pyrifera and Pterygophora californica to irradiances exceeding 900 μmol photons m-2s-1 PAR, which would be expected in very shallow water, for as little as 1 to 4 minutes killed them. However, these irradiances are likely far higher than those that induce photoinhibition in these gametophytes over longer time periods. For instance, Fain and Murray (1982) found that photosynthesis in Macrocystis pyrifera gametophytes is light saturated at 70 μmol photons m-2s-1 PAR, and then begins to decline at 140 μmol photons m-2s-1 PAR. However, this too varies among kelp species, with some deeper water species being far more sensitive to high light. For example, Fejtek et al. (2011) exposed the microscopic gametophytes of Macrocystis pyrifera and Pelagophyucus porra to light levels of 18-20 μmol photons m-2s-1 PAR, which are characteristic of levels observed near the benthos within the shallower (10-15 m) Macrocystis pyrifera forest in Point Loma, California, and found that while Macrocystis pyrifera gametophytes survived and grew well, 100% of the Pelagophycus porra gametophytes died within 24 hrs. In contrast, exposure of the gametophytes of both species to irradiances of 2-4 μmol photons m-2s-1 PAR, which are characteristic of the levels observed near the benthos within the deeper (25-35 m) Pelagophycus porra forest just offshore of the Macrocystis pyrifera forest resulted in high survival and healthy photosystems in both species. Examination of these gametophytes using PAM fluorometry indicated that Pelagophycus porra gametophytes were not able to acclimate to the higher irradiances found within the Macrocystis pyrifera forest. This suggests that, like Macrocystis pyrifera, the shallow depth distribution of Pelagophycus porra is set by the low tolerance of its microscopic stages to high light (Fejtek et al., 2011). Conversely, Vadas (1972) found that gametophytes of Nereocystis leutkeana did not mature when grown at irradiances at or below 4 μmol photons m-2s-1 PAR (161 lux) and concluded that the species’ deep water depth limit was set by low light conditions. This may be related to the fact that Nereocystis leutkeana produces its spores in sori near the surface of the water column where irradiancies are high, or due to the fact that Nereocystis leutkeana occurs in much shallower water than Palagophycus porra. Whether light sensitivity of the microscopic stages is important to setting the distributional limits of even deeper water kelps, such as Pleurophycus gardneri, which grows to greater than 30 m off the central California coast Spalding et al. (2003), Agarum cribrosum, which grows to 40 m off the Gulf of Maine (Vadas and Steneck, 1988), or Laminaria rodriguezii, which grows to >70 m in the Mediterranean Sea, and possibly to 260 m depth in the Adriatic Sea (reviewed in Žuljević et al., 2016), is unknown but certainly worth investigating.

Microscopic Stages Embedded in Communities

Kelp gametophytes occur in variety of microhabitats, including within and on the surface of other algae (Garbary et al., 1999; Hubbard et al., 2004), and on the shells of gastropods (Henríquez et al., 2011). For example, Garbary et al. (1999) observed kelp gametophytes living endophytically in the cell walls of 17 different species of red algae in the San Juan Islands. While they did not identify the gametophyte species, they note that these were within sites that were dominated by several kelps species, namely Alaria marginata, Costaria costata, Laminaria groenlandica, Nereocystis leutkeana, and Agarum fimbriatum. These gametophytes, of which there were hundreds per host individual, exhibited oogonia on raised stocks that extended above the surface of their host. However, these gametophytes may not grow and develop as fast as those on the surface of their host (Hubbard et al., 2004), and not all kelp species (e.g. Nereocystis leutkeana) became endophytic, thus reflecting possible host specificity. Similarly, Desmarestia sp. gametophytes were first discovered in nature embedded in the tissues of the sea pen Ptilosarcus gurneyi by Dube and Ball (1971), which when cultured in the laboratory produced sporophytes. Whether this is due to selective settlement (discussed earlier) or simply a result of random settlement of the zoospores, is unclear, but selection of microhabitats that may offer some protection from benthic grazers has been observed (e.g. Amsler et al., 1992). Indeed, kelp gametophytes are susceptible to grazing by urchins (Dean et al., 1984; Dean et al., 1988), sea stars (Leonard, 1994), and gastropods (Henríquez et al., 2011; Zacher et al., 2016). For example, Henríquez et al. (2011) found that successful recruitment of Macrocystis pyrifera sporophytes along parts of the Chilean coast was linked to the capacity of its gametophytes to colonize secondary substrates, such as those created by the shells of slipper limpets (Crepipatella fecunda), which offered protection from other grazing gastropods. Further, Dean et al. (1988) showed that whit urchins (Lytechinus anamesus) grazed on kelp gametophytes more intensely than other brown algae that compete with the kelps in a southern California kelp forest (e.g. Stephanocystis osmundacea), and therefore may result in exclusion of the kelps and altered community composition. However, climate change may lead to complex ways in which grazers affect kelp microscopic stages. For example, Zacher et al. (2016) found that grazing on the microscopic stages of Alaria esculanta, Laminaria digitata, and Saccharina latissima by limpets was altered under elevated temperatures and levels of sedimentation in a species-specific way, and that this will be important to shaping Arctic kelp communities as the ocean climate changes. This may have significant consequences in the future if climate change also leads to changes in benthic algal compositions, such as the replacement of kelp forests by turf algal communities, which may further alter grazer activities and abundances (Filbee-Dexter and Wernberg, 2018; Zarco-Perello et al., 2021). To better resolve this will require a more detailed examination of available microhabitats in relation to zoospore settlement rates, which may benefit from further advances in eDNA metabarcoding techniques.

Microscopic Stages as a Tool for Conservation and Restoration

As noted earlier in this review, kelps have been experiencing broad scale declines worldwide due to a variety of stressors. This has raised concern about the proper functioning of ecosystems and opened new lines of research into ways to restore them. In particular, research focused on enhancing the abundance and survival of kelp microscopic stages in the field may provide meaningful solutions to kelp restoration efforts (e.g. Carney et al., 2005). Further, it may be possible to identify genetic strains of kelp gametophytes (i.e. ecotypes) that are more resistant to climate change stressors (e.g. Martins et al., 2019) or that maintain genetic diversity of local populations (Barrento et al., 2016), and then delay their development in the laboratory for extended periods so that they can be used as seed stock to replenish kelp populations in areas where they have been lost. This will require more research into how they respond to environmental conditions, what induces them to delay their development, and what causes them to resume development. One promising approach in particular, is the use of “Green Gravel” (small rocks that are inoculated with kelp microscopic stages and then cultured in the laboratory until the kelps are 2-3 cm tall) (Fredriksen et al., 2020). This is now being widely applied along the coast of northern California to help restore the Nereocystis leutkeana populations that have been lost in recent years. It is equivocal, however, whether it is necessary to culture these in the laboratory until the kelps are macroscopic or if the green gravel can be released when the kelps are still microscopic (reviewed in Morris et al., 2020). The latter may reduce costs associated with culturing, energy requirements, and person power, and thus should be considered. This will undoubtedly benefit from further studies that provide a deeper understanding of how these microscopic stages survive, develop and reproduce. Although much has been learned about the ecology of kelp microscopic stages during the last few decades, it is clear that a better understanding of how these cryptic life stages survive and reproduce under different environmental conditions, and how this varies among species is essential if we are to conserve these iconic marine forests in the future.

Concluding Remarks

Kelp forest have been under threat worldwide from numerous environmental factors, including storms, coastal development, urchin grazing, marine heat waves, ENSOs, and climate change. To date, most of the research on how these stressors affect kelps has focused on the large sporophyte life stage that forms these iconic forests. This information has unfortunately not been well integrated into the full life cycle of kelps in nature, which has limited our understanding of the processes by which these forests persist and how they will respond to climate shifts and deforestation in the future. Unlocking this black box surrounding these cryptic life stages will shed light on how their biogeographic ranges and local abundances may change in the future. Indeed, advances in experimental methodologies and a growing body of work on kelp microscopic life stages are demonstrating rather convincingly that these early life stages are instrumental to kelp persistence in a variable ocean. In particular, predicted increases in ocean warming, ocean acidification, large-scale climatic events (e.g., ENSO), storm frequencies, irradiance, and UV radiation, and decreases in salinity, ocean pH, and nutrient availability will likely alter patterns of kelp zoospore dispersal and settlement, gametophyte survival and viability, and sporophyte production. The manner in which these stressors affect kelp microscopic stages, however, will likely vary among kelp species, and with interactions among multiple stressors, leading to complex changes in kelp abundances and geographic distributions. An increased understanding of the ecology and physiology of these cryptic life stages, and of the possibility of maintaining seed stocks of resistant genotypes or using Green Gravel may help guide the conservation of our iconic kelp forest in the future. Regardless, one things is clear; as alluded to in the title of this paper, which mirrors my first conference talk on the subject, it really is ‘the little things’ that matter, and a better understanding of them is fundamental to a complete understanding of kelp forest ecology.

Author Contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Conflict of Interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

I would like to thank M. Foster and D. Reed for their early guidance in subtidal research and for introducing me to kelp microscopic life stages. I am grateful to D. Cie, P. Matson, S. Fejtek, P. Shukla, B. Beckley, M. Morris, S. Small, and especially L. Carney who helped advance this research in my laboratory, and who taught me much about the ecophysiology and ecology of these microscopic forms. I thank KY Kim for introducing me to the methods of studying seaweed physiology. I am grateful to T. Wernberg for inviting me to participate in this special topic of opening the black box surrounding what we know of kelp microscopic stages. T. Winquist, S. Penn and three anonymous reviewers kindly reviewed this manuscript. Lastly, I thank the kelps. Long may they thrive in a changing ocean.

References

Akita S., Murasawa H., Kondo H., Takano Y., Kawakami Y., Nagai S., et al. (2020a). DNA Metabarcoding Analysis of Macroalgal Seed Banks on Shell Surface of the Limpet Niveotctura Pallida. Eur. J. Phycol. 55, 467–477. doi: 10.1080/09670262.2020.1750056

CrossRef Full Text | Google Scholar

Akita S., Murasawa H., Tnako Y., Kawakami Y., Fujita D., Nagai S. (2020b). Variation in “Bank of Microscopic Forms” in Urchin Barren Coast: Detection Using DNA Microbarcoding Based on High-Throughput Sequencing. J. Appl. Phycol. 32, 2115–2124. doi: 10.1007/s10811-020-02122-3

CrossRef Full Text | Google Scholar

Akita S., Tanako Y., Nagai S., Kuwahara H., Kajihara R., Tanabe A. S., et al. (2019). Rapid Detection of Macroalgal Seed Bank on Cobbles: Application of DNA Metabarcoding Using Next-Generation Sequencing. J. Appl. Phycol. 31, 2743–2753. doi: 10.1007/s10811-018-1730-9

CrossRef Full Text | Google Scholar

Alvarez-Buylla E. R., Martínez-Ramos M. (1990). Seed Bank Versus Seed Rain in the Regeneration of a Tropical Pioneer Tree. Oecologia 84, 314–325. doi: 10.1007/BF00329755

PubMed Abstract | CrossRef Full Text | Google Scholar

Amsler C. D. (1988). Kelp Spore Photosynthesis, Respiration, and Pigmentation. J. Phycol. 24 (Suppl.), 4.

Google Scholar

Amsler C. D., Neushul M. (1989a). Diel Periodicity of Spore Release From the Kelp Nereocystis Luetkeana (Mertens) Postels Et Ruprecht. J. Exp. Marine Biol. Ecol. 134 (2), 117–127. doi: 10.1016/0022-0981(90)90104-K

CrossRef Full Text | Google Scholar

Amsler C. D., Neushul M. (1989b). Chemotactic Effects of Nutrients on Spores of the Kelps Macrocytis Pyrifera and Pterygophora California. Marine Biol. 102 (4), 557–564. doi: 10.1007/BF00438358

CrossRef Full Text | Google Scholar

Amsler C. D., Neushul M. (1990). Nutrient Stimulation of Spore Settlement in the Kelps Pterygophora Californica and Macrocystis Pyrifera. Marine Biol. 107 (2), 297–304. doi: 10.1007/BF01319829

CrossRef Full Text | Google Scholar

Amsler C. D., Neushul M. (1991). Photosynthetic Physiology and Chemical; Composition of Spores of the Kelps Macrocystis Pyrifera, Nereocystis Leutkeana, Laminaria Farlowii, and Pterygophora Californica (Phaeophyceae). J. Phycol. 16, 617–619. doi: 10.1111/j.1529-8817.1980.tb03080.x

CrossRef Full Text | Google Scholar

Amsler C. D., Reed D. C., Neushul M. (1992). The Microclimate Inhabited by Macroalgal Propagules. Br. Phycol. J. 27, 253–270. doi: 10.1080/00071619200650251

CrossRef Full Text | Google Scholar

Amsler C. D., Searles R. B. (1980). Vertical Distribution of Seaweed Spores in the Water Column Offshore of North Carolina. J. Phycol 16, 617–619. doi: 10.1111/j.1529-8817.1980.tb03080.x

CrossRef Full Text | Google Scholar

Anderson E. K., North W. J. (1966). In Situ Studies of Spore Production and Dispersal in the Giant Kelp, Macrocystis Pyrifera. Proc. Int. Seaweed Symposium 5, 73–86. doi: 10.1016/B978-0-08-011841-3.50011-2

CrossRef Full Text | Google Scholar

Arendt A., Walsh J., Harrison W. (2009). Changes of Glaciers and Climate in Northwestern North America During the Late Twentieth Century. J. Climate 22, 4117–4134. doi: 10.1175/2009JCLI2784.1

CrossRef Full Text | Google Scholar

Armstrong I. A. (1977). “Weather and Climate,” in The Environment of Amchitka Island, Alaska. US Energy Research and Development Administration. Eds. Merritt M. L., Fuller R. G. (US Division of Military) pp53–pp58.

Google Scholar

Assis J., Araujo M. B., Serrao E. A. (2018). Projected Climate Changes Threaten Ancient Refugia of Kelp Forests in the North Atlantic. Global Chang. Biol. 24, e55–e66. doi: 10.1111/gcb.13818

CrossRef Full Text | Google Scholar

Assis J., Berecibar E., Claro B., Alberto F., Reed D., Raimondi P., et al. (2017). Major Shifts at the Range Edge of Marine Forests: The Combined Effects of Climate Changes and Limited Dispersal. Sci. Rep. 7 (1), 1–10. doi: 10.1038/srep44348

PubMed Abstract | CrossRef Full Text | Google Scholar

Baker H. G. (1955). Self-Compatibility and Establishment After 'Long-Distance' Dispersal. Evolution 9 (3), 347–349. doi: 10.2307/2405656

CrossRef Full Text | Google Scholar

Bakun A., Black B. A., Bograd S. J., Garcia-Reyes M., Miller A. J., Rykaczewski R. R., et al. (2015). Anticipated Effects of Climate Change on Coastal Upwelling Ecosystems. Curr. Climate Change Rep. 1 (2), 85–93. doi: 10.1007/s40641-015-0008-4

CrossRef Full Text | Google Scholar

Barner A. K., Pfister C. A., Wootton J. T. (2010). The Mixed Mating System of the Sea Palm Kelp Postelsia Palmaeformis: Few Costs to Selfing. Proc. R. Soc. B: Biol. Sci. 278 (1710), 1347–1355. doi: 10.1098/rspb.2010.1928

CrossRef Full Text | Google Scholar

Barrento S., Camus C., Sousa-Pinto I., Buschmann A. H. (2016). Germplasm Banking of the Giant Kelp: Our Biological Insurance in a Changing Environment. Algal Res. 13, 134–140. doi: 10.1016/j.algal.2015.11.024

CrossRef Full Text | Google Scholar

Batista M. B., Anderson A. B., Sanches P. F., Polito P. S., Silveira T. C. L., Velez-Rubio G. M., et al. (2018). Kelps’ Long-Distance Dispersal: Role of Ecological/Oceanographic Processes and Implications to Marine Forest Conservation. Diversity 10, 11. doi: 10.3390/d10010011

CrossRef Full Text | Google Scholar

Beas R., Micheli F., Woodson C., Carr M., Malone D., Torre J., et al. (2020). Geographic Variation in Responses of Kelp Forest Communities to Recent Climatic Changes. Global Change Biol. 26, 6467–6473. doi: 10.1111/gcb.15273

CrossRef Full Text | Google Scholar

Beckley B. A., Edwards M. S. (2021). Mechanisms Leading to Recruitment Inhibition of Macrocystis Pyrifera by an Understory Alga. Marine Ecol. Prog. Ser. 657, 59–71. doi: 10.3354/meps13550

CrossRef Full Text | Google Scholar

Behrenfeld M. J., O’Malley R. T., Siegel D. A., McClain C. R., Sarmiento J. L., Feldman G. C., et al. (2006). Climate-Driven Trends in Contemporary Ocean Productivity. Nature 444 (7120), 752–755. doi: 10.1038/nature05317

PubMed Abstract | CrossRef Full Text | Google Scholar

Bennett S., Wernberg T., de Bettignies T., Kendrick G. A., Anderson R. J., Bolton J. J., et al. (2015). Canopy Interactions and Physical Stress Gradients in Subtidal Communities. Ecol. Lett. 18, 636–645. doi: 10.1111/ele.12446

PubMed Abstract | CrossRef Full Text | Google Scholar

Bews E., Booher L., Polizzi T., Long C., Kim J.-H., Edwards M. S. (2021). Effects of Salinity and Nutrients on Metabolism and Growth of Ulva Lactuca: Implications for Bioremediation of Coastal Watersheds. Marine Pollution Bull. 166, 121299. doi: 10.1016/j.marpolbul.2021.112199

CrossRef Full Text | Google Scholar

Bieniek P. A., Walsh J. E., Thoman R. L., Bhatt U. S. (2014). Using Climate Divisions to Analyze Variations and Trends in Alaska Temperature and Precipitation. J. Climate 27, 2800–2818. doi: 10.1175/JCLI-D-13-00342.1

CrossRef Full Text | Google Scholar

Bischof K., Gomez I., Molis M., Hanelt D., Karsten U., Lüder U., et al. (2006). Ultraviolet Radiation Shapes Seaweed Communities. Rev. Environ. Sci. Bio/Technol 5 (2), 141–166. doi: 10.1007/s11157-006-0002-3

CrossRef Full Text | Google Scholar

Blain C. O., Hansen S. C., Shears N. T. (2021). Coastal Darkening Substantially Limits the Contribution of Kelp to Coastal Carbon Cycles. Global Change Biol. 27 (21), 5547–5563. doi: 10.1111/gcb.15837

CrossRef Full Text | Google Scholar

Blanchette C. A. (1996). Seasonal Patterns of Disturbance Influence Recruitment of the Sea Palm, Postelsia Palmaeformis. J. Exp. Marine Biol. Ecol. 197, 1–14. doi: 10.1016/0022-0981(95)00141-7

CrossRef Full Text | Google Scholar

Bolton J. J., Lüning K. (1982). Optimal Growth and Maximal Survival Temperatures of Atlantic Laminaria Species (Phaeophyta) in Culture. Marine Biol. 66, 89–94. doi: 10.1007/BF00397259

CrossRef Full Text | Google Scholar

Borras-Chavez R., Edwards M. S., Alvizu-Huiguera D. L., Montesinos E., Hernandez-Carmona G., Breceño-Domínguez D. (2016). Repetitive Harvesting of Macrocystis Pyrifera and Its Effects on Chemical Constitutes of Economic Value. Botanica Marina 59, 63–71. doi: 10.1515/bot-2015-0028

CrossRef Full Text | Google Scholar

Borras-Chavez R., Edwards M. S., Vasquez J. (2012). Testing Sustainable Management in Northern Chile: Harvesting Macrocystis Pyrifera (Phaeophyceae, Laminariales). A Case Study. J. Appl. Phycol. 24, 1655–1665. doi: 10.1007/s10811-012-9829-x

CrossRef Full Text | Google Scholar

Boyer T. P., Levitus S., Antonov J., Locarnini R., Garcia H. (2005). Linear Trends in Salinity for the World Ocean 1955–1998. Geophys Res. Lett. 32, L01604. doi: 10.1029/2004GL021791

CrossRef Full Text | Google Scholar

Brown M. B., Edwards M. S., Kim K. Y. (2014). Effects of Climate Change on the Physiology of Giant Kelp, Macrocystis Pyrifera, and Grazing by Purple Urchin, Strongylocentrotus Purpuratus. Algae 29 (3), 203–215. doi: 10.4490/algae.2014.29.3.203

CrossRef Full Text | Google Scholar

Brzezinski M., Reed D. C., Amsler C. D. (1993). Neutral Lipids as Major Storage Products in Zoospores of the Giant Kelp, Macrocystis Pyrifera. J. Phycol. 29, 16–23. doi: 10.1111/j.1529-8817.1993.tb00275.x

CrossRef Full Text | Google Scholar

Burgman M. A., Gerard V. A. (1990). A Stage-Structured, Stochastic Population Model for the Giant Kelp Macrocystis Pyrifera. Marine Biol. 105, 15–23. doi: 10.1007/BF01344266

CrossRef Full Text | Google Scholar

Buschmann A. H., Vásquez J. A., Osorio P., Reyes E., Filún L., Hernández-González M. C., et al. (2004). The Effect of Water Movement, Temperature and Salinity on Abundance and Reproductive Patterns of Macrocystis Spp. (Phaeophyta) at Different Latitudes in Chile. Marine Biol. 145 (5), 849–862. doi: 10.1007/s00227-004-1393-8

CrossRef Full Text | Google Scholar

Byrnes J. E., Reed D. C., Cardniale B. J., Cavanaugh K. C., Holbrook S. J., Schmitt R. J. (2011). Climate-Driven Increases in Storm Frequency Simplify Kelp Forest Food Webs. Global Change Biol. 17 (8), 2513–2524. doi: 10.1111/j.1365-2486.2011.02409.x

CrossRef Full Text | Google Scholar

Camus C., Solas M., Martínez C., Vargas J., Garcés C., Gil-Kodaka P., et al. (2021). Mates Matter: Gametophyte Kinship Recognition and Inbreeding in the Giant Kelp, Macrocystis Pyrifera (Laminariales, Phaeophyceae). J. Phycol. 57 (3), 711–725. doi: 10.1111/jpy.13146

PubMed Abstract | CrossRef Full Text | Google Scholar

Careau V., Gifford M. E., Biro P. A. (2014). Individual (Co)Variation in Thermal Reaction Norms of Standard and Maximal Metabolic Rates in Wild-Caught Slimy Salamanders. Funct. Ecol. 28, 1175–1118. doi: 10.1111/1365-2435.12259

CrossRef Full Text | Google Scholar

Carney L. T. (2011). A Multispecies Laboratory Assessment of Rapid Sporophyte Recruitment From Delayed Kelp Gametophytes. J. Phycol 47, 244–251. doi: 10.1111/j.1529-8817.2011.00957.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Carney L. T., Bohonak A. J., Edwards M. S., Alberto F. (2013). Genetic and Experimental Evidence for a Mixed-Age, Mixed-Origin Bank of Kelp Microscopic Stages in Southern California. Ecology 94, 1955–1965. doi: 10.1890/13-0250.1

PubMed Abstract | CrossRef Full Text | Google Scholar

Carney L. T., Edwards M. S. (2006). Cryptic Processes in the Sea: A Review of Delayed Development in the Microscopic Life Stages of Marine Macroalgae. Algae 21, 161–168. doi: 10.4490/ALGAE.2006.21.2.161

CrossRef Full Text | Google Scholar

Carney L. T., Edwards M. S. (2010). Role of Nutrient Fluctuations and Delayed Development in Gametophyte Reproduction by Macrocystis Pyrifera (Phaeophyceae) in Southern California. J. Phycol. 46, 987–996. doi: 10.1111/j.1529-8817.2010.00882.x

CrossRef Full Text | Google Scholar

Carney L. T., Waaland J. R., Klinger T., Ewing K. (2005). Restoration of the Bull Kelp Nereocystis Leutkeana in Nearshore Rocky Habitats. Marine Ecol. Prog. Ser. 302, 49–61. doi: 10.3354/meps302049

CrossRef Full Text | Google Scholar

Cavalcanti G., Shukla P., Morris M., Ribeiro B., Foley M., Doane M., et al. (2018). Rhodolith Holobionts in a Changing Ocean: Host-Microbes Interactions Mediate Coralline Algae Resilience Under Ocean Acidification. BMC Genomics 19, 701–714. doi: 10.1186/s12864-018-5064-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Cavanaugh K. C., Reed D. C., Bell T. W., Castorani M. C., Beas-Luna R. (2019). Spatial Variability in the Resistance and Resilience of Giant Kelp in Southern and Baja California to a Multiyear Heatwave. Front. Marine Sci. 6, 413. doi: 10.3389/fmars.2019.00413

CrossRef Full Text | Google Scholar

Cavanaugh K. C., Siegel D. A., Reed D. C., Dennison P. E. (2011). Environmental Controls of Giant-Kelp Biomass in the Santa Barbara Channel, California. Marine Ecol. Prog. Ser. 429, 1–17. doi: 10.3354/meps09141

CrossRef Full Text | Google Scholar

Cetina-Heredia P., Roughan M., van Sebille E., Feng M., Coleman M. A. (2015). Strengthened Currents Override the Effect of Warming on Lobster Larval Dispersal and Survival. Global Change Biol. 21, 4377–4386. doi: 10.1111/gcb.13063

CrossRef Full Text | Google Scholar

Chapman A. R. O. (1986). Population and Community Ecology of Seaweeds. Adv. Marine Biol. 23, 1–161. doi: 10.1016/S0065-2881(08)60108-X

CrossRef Full Text | Google Scholar

Cie D. K., Edwards M. S. (2008). The Effects of High Irradiance on the Settlement Competency and Viability of Kelp Zoospores. J. Phycol. 44, 495–500. doi: 10.1111/j.1529-8817.2008.00464.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Cie D., Edwards M. (2011). Vertical Distribution of Kelp Zoospores. Phycologia 50, 340–350. doi: 10.2216/10-48.1

CrossRef Full Text | Google Scholar

Clark R. P., Edwards M. S., Foster M. S. (2004). Effects of Shade From Multiple Kelp Canopies on an Understory Algal Assemblage. Marine Ecol. Prog. Ser. 267, 107–119. doi: 10.3354/meps267107

CrossRef Full Text | Google Scholar

Cohen D., Levin S. A. (1987). “The Interaction Between Dispersal and Dormancy Strategies in Varying and Heterogeneous Environments,” in Mathematical Topics in Population Biology, Morphogenesis and Neurosciences (Springer, Berlin, Heidelberg), 110–122.

Google Scholar

Cole K. (1964). Induced Fluorescence in Gametophytes in Some Laminariales. Can. J. Bot. 42, 1173–1183. doi: 10.1139/b64-113

CrossRef Full Text | Google Scholar

Connell S. D., Russell B. D. (2010). The Direst Effects of Increasing CO2 and Temperature on Non-Calcifying Organisms: Increasing the Potential for Phase Shifts in Kelp Forests. Proc. R. Soc. B. 277, 1409–1415. doi: 10.1098/rspb.2009.2069

CrossRef Full Text | Google Scholar

Connell S. D., Russell B. D., Turner D. J., Shepherd S. A., Kildea T., Miller D., et al. (2008). Recovering a Lost Baseline: Missing Kelp Forests From a Metropolitan Coast. Marine Ecol. Prog. Series. 360, 63–72. doi: 10.3354/meps07526

CrossRef Full Text | Google Scholar

Corrano M. W., Corrano C. J., Edwards M. S., Al-Adilah H., Fontana Y., Sayer M. D. J., et al. (2021). Laminaria Kelps Impact Iodine Speciation Chemistry in Coastal Seawater. Estuar Coastal Shelf Sci. 262, 107531. doi: 10.1016/j.ecss.2021.107531

CrossRef Full Text | Google Scholar

Corrano M. W., Yarimizu K., Gonzales J. L., Edwards M. S., Tymon T. M., Küpper F. C., et al. (2020). The Influence of Marine Algae on Iodine Speciation in the Coastal Ocean. Algae 35, 1–10. doi: 10.4490/algae.2020.35.5.25

CrossRef Full Text | Google Scholar

Coyer J. (1984). The Invertebrate Assemblage Associated With the Giant Kelp, Macrocystis Pyrifera, at Santa Catalina Island, California: A General Description With Emphasis on Amphipods, Copepods, Mysids, and Shrimps. Fish Bull. 82, 55–66.

Google Scholar

Coyer J. A., Olsen J. L., Stam W. T. (1997). Genetic Variability and Spatial Separation in the Sea Palm Kelp Postelsia Palmaeformis (Phaeophyceae) as Assessed With M13 Fingerprints and RAPDs. J. Phycol. 33, 561–568. doi: 10.1111/j.0022-3646.1997.00561.x

CrossRef Full Text | Google Scholar

Crowe J. H. (1971). Anhydrobiosis: An Unsolved Problem. Am. Nat. 105, 563–573. doi: 10.1086/282745

CrossRef Full Text | Google Scholar

Crow J. F., Kimura M. (1965). Evolution in Sexual and Asexual Populations. Am. Nat. 99, 515–518. doi: 10.1086/282389

CrossRef Full Text | Google Scholar

Dam H. G., Baumann H. (2017). Climate Change, Zooplankton and Fisheries. Climate Change Impacts Fish Aquacult: A Global Anal. 2, 851–874. doi: 10.1002/9781119154051.ch25

CrossRef Full Text | Google Scholar

Dayton P. (1973). Dispersion and Persistence of the Annual Intertidal Alga Postelsia Palmaeformis. Ecology 54, 433–438. doi: 10.2307/1934353

CrossRef Full Text | Google Scholar

Dayton P. K. (1985). Ecology of Kelp Communities. Annu. Rev. Syst. Ecol. 16, 215–245. doi: 10.1146/annurev.es.16.110185.001243

CrossRef Full Text | Google Scholar

Dayton P. K., Currie V., Gerrodette T., Keller B. D., Rosenthal R., Ven Tresca D. (1984). Patch Dynamics and Stability of Some California Kelp Communities. Ecol. Monogr. 54, 253–289. doi: 10.2307/1942498

CrossRef Full Text | Google Scholar

Dean T. A., Jacobsen F. R., Thies K., Lagos S. L. (1988). Differential Effects of Grazing by White Sea Urchins on Recruitment of Brown Algae. Marine Ecol. Prog. Ser. 48 (1), 99–102. doi: 10.3354/meps048099

CrossRef Full Text | Google Scholar

Dean T. A., Schroeter S. C., Dixon J. D. (1984). Effects of Grazing by Two Species of Sea Urchins (Strongylocentrotus Franciscanus and Lytechinus Anamesus) on Recruitment and Survival of Two Species of Kelp (Macrocystis Pyrifera and Pterygophora Californica). Mar. Biol. 78, 301–313. doi: 10.1007/BF00393016

CrossRef Full Text | Google Scholar

Deiman M., Iken K., Konar B. (2012). Susceptibility of Nereocystis Leutkeana (Laninariales, Ochrophyta) and Eualaria Fistulosa (Laminariales, Ochrophyta) Spores to Sedimentation. Algae 27, 115–123. doi: 10.4490/algae.2012.27.2.115

CrossRef Full Text | Google Scholar

Destombe C., Valeria Oppliger L. (2011). Male Gametophyte Fragmentation in Laminaria digitata: A Life History Strategy to Enhance Reproductive Success. CBM-Cah. Biol. Mar. 52 (4), 385.

Google Scholar

Devinny J. S., Volse L. A. (1978). Effects of Sediments on the Development of Macrocystis Pyrifera Gametophytes. Marine Biol. 48, 343–348. doi: 10.1007/BF00391638

CrossRef Full Text | Google Scholar

Deysher L. E., Dean T. A. (1984). Critical Irradiance Levels and the Interactive Effects of Quantum Irradiance and Dose on Gametogenesis in the Giant-Kelp, Macrocystis Pyrifera. J. Phycol 20, 520–524. doi: 10.1111/j.0022-3646.1984.00520.x

CrossRef Full Text | Google Scholar

Deysher L. E., Dean T. A. (1986). In Situ Recruitment of Sporophytes of the Giant-Kelp, Macrocystis-Pyrifera (L) Agardh,C.A. - Effects of Physical Factors. J. Exp. Marine Biol. Ecol. 103, 41–63. doi: 10.1016/0022-0981(86)90131-0

CrossRef Full Text | Google Scholar

Dolinar D., Edwards M. (2021). The Zombification and Reanimation of Purple Urchins (Strongylocentrotus Purpuratus) in Response to Macroalgal Availability. J. Exp. Marine Biol. Ecol. 545, 151646. doi: 10.1016/j.jembe.2021.151646

CrossRef Full Text | Google Scholar

Doney S. C., Fabry V. J., Feely R. A., Kleypas J. A. (2009). Ocean Acidification: The Other CO2 Problem. Annu. Rev. Marine Sci. 1, 169–192. doi: 10.1146/annurev.marine.010908.163834

CrossRef Full Text | Google Scholar

Doney S. C., Ruckelshaus M., Duffy J. E., Barry J. P., Chan F., English C. M., et al. (2012). Climate Change Impacts on Marine Ecosystems. Annu. Rev. Marine Sci. 4, 11–37. doi: 10.1146/annurev-marine-041911-111611

CrossRef Full Text | Google Scholar

Dring M. J. (1988). Photocontrol of Development in Algae. Annu. Rev. Plant Physiol. Plant Mol. Biol. 39, 157–174. doi: 10.1146/annurev.pp.39.060188.001105

CrossRef Full Text | Google Scholar

Dube M. A., Ball E. (1971). Desmarestia Sp. Associated With the Sea Pen Ptilosarcus Gurney (Gray). J. Phycol 7 (3), 218–220. doi: 10.1111/j.1529-8817.1971.tb01506.x

CrossRef Full Text | Google Scholar

Dyurgerov M. B., Meier M. F. (2005). Glaciers and the Changing Earth: A 2004 Snapshot. INSTAAR Univ. Colorado at Boulder Occasional Paper 58, 1–117.

Google Scholar

Ebbing A., Pierik R., Bouma T., Kromkamp J. C., Timmermans K. (2020). How Light and Biomass Density Influence the Reproduction of Delayed Saccharina Latissima Gametophytes (Phaeophyceae). J. Phycol. 56 (3), 709–718. doi: 10.1111/jpy.12976

PubMed Abstract | CrossRef Full Text | Google Scholar

Ebeling A. W., Laur D. R., Rowley R. J. (1985). Severe Storm Disturbances and Reversal of Community Structure in a Southern California Kelp Forest. Marine Biol. 84, 287–294. doi: 10.1007/BF00392498

CrossRef Full Text | Google Scholar

Edwards M. S. (1998). Effects of Long-Term Kelp Canopy Exclusion on the Abundance of the Annual Alga Desmarestia Ligulata. J. Exp. Marine Biol. Ecol. 228, 309–326. doi: 10.1016/S0022-0981(98)00046-X

CrossRef Full Text | Google Scholar

Edwards M. S. (1999) in Proceedings form the 16th Annual International Seaweeds Symposium. Hydrobiologia (Springer, Dordrecht) Vol. 398/399. 253–259.

Google Scholar

Edwards M. S. (2000). The Role of Alternate Life-History Stages of a Marine Macroalga: A Seed Bank Analogue? Ecology 8, 2402–2415. doi: 10.1890/0012-9658(2000)081[2404:TROALH]2.0.CO;2

CrossRef Full Text | Google Scholar

Edwards M. S. (2004). Estimating Scale Dependency in Disturbance Impacts: El Niños and Giant Kelp Forests in the Northeast Pacific. Oecologia 138, 436–447. doi: 10.1007/s00442-003-1452-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Edwards M. S. (2019). Comparing the Impacts of Four ENSO Events on Giant Kelp (Macrocystis Pyrifera) in the Northeast Pacific Ocean. Algae. 34, 141–151. doi: 10.4490/algae.2019.34.5.4

CrossRef Full Text | Google Scholar

Edwards M. S., Estes J. A. (2006). Catastrophe, Recovery, and Range Limitation in NE Pacific Kelp Forests: A Large-Scale Perspective. Marine Ecol. Prog. Ser. 320, 79–87. doi: 10.3354/meps320079

CrossRef Full Text | Google Scholar

Edwards M. S., Hernández-Carmona G. (2005). Delayed Recovery of Giant Kelp Near Its Southern Range Limit in the North Pacific Following El Niño. Marine Biol. 147, 273–279. doi: 10.1007/s00227-004-1548-7

CrossRef Full Text | Google Scholar

Edwards M. S., Kim K. I. Y. (2010). Diurnal Variation in Photosynthetic Performance in Giant Kelp Macrocystis Pyrifera (Phaeophyceae, Laminariales) at Different Depths as Estimated Using PAM Fluorometry. Aquat. Bot. 92, 119–128. doi: 10.1016/j.aquabot.2009.10.017

CrossRef Full Text | Google Scholar

Edwards M. S., Konar B. K. (2012). A Comparison of Dragon Kelp, Eualaria Fistulosa, (Phaeophyceae) Fecundity in Urchin Barrens and Nearby Kelp Beds Throughout the Aleutian Archipelago. J. Phycol 48, 897–901. doi: 10.1111/j.1529-8817.2012.01139.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Edwards M. S., Konar B. K., Kim J. H., Gabara S., Sullaway G., McHugh T. A., et al. (2020). Marine Deforestation Leads to Widespread Loss of Ecosystem Function. PloS One 15 (3), e0226173. doi: 10.1371/journal.pone.0226173

PubMed Abstract | CrossRef Full Text | Google Scholar

Estes J. A., Tinker M. T., Williams T. M., Doak D. F. (1998). Killer Whale Predation on Sea Otters Linking Oceanic and Nearshore Ecosystems. Science 282 (5388), 473–476. doi: 10.1126/science.282.5388.473

PubMed Abstract | CrossRef Full Text | Google Scholar

Evans L., Edwards M. S. (2011). Bioaccumulation of Copper and Zinc by the Giant Kelp, Macrocystis Pyrifera. Algae 26, 265–275. doi: 10.4490/algae.2011.26.3.265

CrossRef Full Text | Google Scholar

Fagerli C. W., Norderhaug K. M., Christie H. C. (2013). Lack of Sea Urchin Settlement May Explain Kelp Forest Recovery in Overgrazed Areas in Norway. Marine Ecol. Prog. Ser. 488, 119–132. doi: 10.3354/meps10413

CrossRef Full Text | Google Scholar

Fain S. R., Murray S. N. (1982). Effects of Light and Temperature on Net Photosynthesis and Dark Respiration of Gametophytes and Embryonic Sporophytes of Macrocystis Pyrifera. J. Phycol 18, 92–98. doi: 10.1111/j.1529-8817.1982.tb03161.x

CrossRef Full Text | Google Scholar

Fejtek S. M., Edwards M. S., Kim K. Y. (2011). Elk Kelp, Pelagophycus Porra, Distribution Limited Due to Susceptibility of Microscopic Stages to High Light. J. Exp. Marine Biol. Ecol. 396, 194–201. doi: 10.1016/j.jembe.2010.10.022

CrossRef Full Text | Google Scholar

Filbee-Dexter K., Feehan C. J., Scheibling R. E. (2016). Large-Scale Degradation of a Kelp Ecosystem in an Ocean Warming Hotspot. Marine Ecol. Prog. Ser. 543, 141–152. doi: 10.3354/meps11554

CrossRef Full Text | Google Scholar

Filbee-Dexter K., Wernberg T. (2018). Rise of Turfs: A New Battlefront for Globally Declining Kelp Forests. Bioscience 68 (2), 64–76. doi: 10.1093/biosci/bix147

CrossRef Full Text | Google Scholar

Foden J., Devlin M. J., Mills D. K., Malcolm S. J. (2010). Searching for Undesirable Disturbance: An Application of the OSPAR Eutrophication Assessment Method to Marine Waters of England and Wales. Biogeochemistry 106, 157–175. doi: 10.1007/s10533-010-9475-9

CrossRef Full Text | Google Scholar

Foster M. S., Schiel D. R. (2010). Loss of Predators and the Collapse of Southern California Kelp Forests (?): Alternatives, Explanations and Generalizations. J. Exp. Marine Biol. Ecol. 393, 59–70. doi: 10.1016/j.jembe.2010.07.002

CrossRef Full Text | Google Scholar

Fox C. H., Swanson A. K. (2007). Nested PCR Detection of Microscopic Life Stages of Laminarian Macroalgae and Comparison With Adult Forms Along Intertidal Height Gradients. Marine Ecol. Prog. Ser. 332, 1–10. doi: 10.3354/meps332001

CrossRef Full Text | Google Scholar

Franke K., Liesner D., Heesch S., Bartsch I. (2021). Looks can be Deceiving: Contrasting Temperature Characteristics of Two Morphologically Similar Kelp Species Co-Occurring in the Arctic. Botanica Marina 64 (3), 163–175. doi: 10.1515/bot-2021-0014

CrossRef Full Text | Google Scholar

Fredersdorf J., Müller R., Becker S., Wiencke C., Bischof K. (2009). Interactive Effects of Radiation, Temperature and Salinity on Different Life History Stages of the Arctic Kelp(Phaeophyceae). Oecologia. 161 (3), 483–492. doi: 10.1007/s00442-009-1326-9

CrossRef Full Text | Google Scholar

Fredriksen S., Filbee-Dexter K., Norderhaug K. M., Steen H., Bodvin T., Coleman M. A., et al. (2020). Green Gravel: A Novel Restoration Tool to Combat Kelp Forest Decline. Sci. Rep. 10 (1), 1–7. doi: 10.1038/s41598-020-60553-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Gabara S., Konar B., Edwards M. (2021). Biodiversity Loss Leads to Reductions in Community-Wide Trophic Complexity. Ecosphere 12 (2), e03361. doi: 10.1002/ecs2.3361

CrossRef Full Text | Google Scholar

Gaitán-Espitia J. D., Hancock J. R., Padilla-Gamiño J. L., Rivest E. B., Blanchette C. A., Reed D. C., et al. (2014). Interactive Effects of Elevated Temperature and Pco2 on Early-Life-History Stages of the Giant Kelp Macrocystis Pyrifera. J. Exp. Marine Biol. Ecol. 457, 51–58. doi: 10.1016/j.jembe.2014.03.018

CrossRef Full Text | Google Scholar

Garbary D. J., Kim K. Y., Klinger T., Duggins D. (1999). Red Algae as Hosts for Endophytic Kelp Gametophytes. Marine Biol. 135 (1), 35–40. doi: 10.1007/s002270050598

CrossRef Full Text | Google Scholar

Gaylord B., Reed D. C., Raimondi P. T., Washburn L. (2006). Macroalgal Spore Dispersal in Coastal Environments: Mechanistic Insights Revealed by Theory and Experiment. Ecol. Monogr. 76, 481–502. doi: 10.1890/0012-9615(2006)076[0481:MSDICE]2.0.CO;2

CrossRef Full Text | Google Scholar

Gaylord B., Reed D. C., Raimondi P. T., Washburn L., McLean S. R. (2002). A Physically Based Model of Macroalgal Macroalgal Spore Dispersal in the Wave and Current-Dominated Nearshore. Ecology 83, 1239–1251. doi: 10.1890/0012-9658(2002)083[1239:APBMOM]2.0.CO;2

CrossRef Full Text | Google Scholar

Geiser F. (2004). Metabolic Rate and Body Temperature Reduction During Hibernation and Daily Torpor. Annu. Rev. Physiol. 66, 239–274. doi: 10.1146/annurev.physiol.66.032102.115105

PubMed Abstract | CrossRef Full Text | Google Scholar

Gerard V. A. (1988). Ecotypic Differentiation in Light-Related Traits of the Kelp Laminaria Saccharina. Marine Biol. 97 (1), 25–36. doi: 10.1007/BF00391242

CrossRef Full Text | Google Scholar

Gerard V. A. (1990). Ecotypic Differentiation in the Kelp Laminaria Saccharina—Phase Specific Adaptation in a Complex Life Cycle. Marine Biol. 107, 519–528. doi: 10.1007/BF01313437

CrossRef Full Text | Google Scholar

Gonzalez J. G., Tymon T., Küpper F. C., Edwards M. S., Corrano C. J. (2017). The Potential Role of Kelp Forests on Iodine Speciation in Coastal Seawater. PloS One 12 (12), e0189559. doi: 10.1371/journal.pone.0189559

PubMed Abstract | CrossRef Full Text | Google Scholar

Graham M. H. (1996). Effect of High Irradiance on Recruitment of the Giant Kelp Macrocystis (Phaeophyta) in Shallow Water. J. Phycol 32 (6), 903–906. doi: 10.1111/j.0022-3646.1996.00903.x

CrossRef Full Text | Google Scholar

Graham M. H. (1997). Factors Determining the Upper Limit of Giant Kelp Macrocystis Pyrifera Agardh, Along the Monterey Peninsula, Central California, USA. J. Exp. Marine Biol. Ecol. 218, 127–149. doi: 10.1016/S0022-0981(97)00072-5

CrossRef Full Text | Google Scholar

Graham M. H. (1999). Identification of Kelp Zoospores From in Situ Plankton Samples. Marine Biol. 135, 709–720. doi: 10.1007/s002270050672

CrossRef Full Text | Google Scholar

Graham M. H. (2003). Coupling Propagule Output to Supply at the Edge and Interior of a Giant Kelp Forest. Ecology 84 (5), 1250–1264. doi: 10.1890/0012-9658(2003)084[1250:CPOTSA]2.0.CO;2

CrossRef Full Text | Google Scholar

Graham M. H., Mitchell B. G. (1999). Obtaining Absorption Spectra From Individual Macroalgal Spores Using Microphotometry. Hydrobiologia 398/399, 231–239. doi: 10.1023/A:1017009411367

CrossRef Full Text | Google Scholar

Graham M. H., Vasquez J. A., Buschmann A. H. (2007). Global Ecology of the Giant Kelp Macrocystis: From Ecotypes to Ecosystems. Oceanogr. Marine Biol. 45, 39.

Google Scholar

Gregg W. W., Casey N. W., McClain C. R. (2005). Recent Trends in Global Ocean Chlorophyll. Geophys Res. Lett. 32, L03606. doi: 10.1029/2004GL021808

CrossRef Full Text | Google Scholar

Grice G. D., Marcus N. H. (1981). Dormant Eggs of Marine Copepods. Oceanogr Marine Biol. Annu. Review. 19, 125–140.

Google Scholar

Guidetti R., Altiero T., Rebecch L. (2011). On Dormancy Strategies in Tartigrades. J. Insect Physiol. 57, 567–576. doi: 10.1016/j.jinsphys.2011.03.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Harley C. D., Anderson K. M., Demes K. W., Jorve J. P., Kordas R. L., Coyle T. A., et al. (2012). Effects of Climate Change on Global Seaweed Communities. J. Phycol. 48 (5), 1064–1078. doi: 10.1111/j.1529-8817.2012.01224.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hays G. C. (2017). Ocean Currents and Marine Life. Curr. Biol. 27, R431–R510. doi: 10.1016/j.cub.2017.01.044

PubMed Abstract | CrossRef Full Text | Google Scholar

Heldmaier G., Ortmann S., Elvert R. (2004). Natural Hypometabolism During Hibernation and Daily Torpor in Mammals. Respiration Physiol. Neurobiol. 141, 317–329. doi: 10.1016/j.resp.2004.03.014

CrossRef Full Text | Google Scholar

Henkel S. K., Hofmann G. E. (2008). Thermal Ecophysiology of Gametophytes Cultured From Invasive Undaria Pinnatifida (Harvey) Suringar in Coastal California Harbors. J. Exp. Marine Biol. Ecol. 367 (2), 164–173. doi: 10.1016/j.jembe.2008.09.010

CrossRef Full Text | Google Scholar

Henríquez L. A., Buschmann A. H., Maldonado M. A., Graham M. H. (2011). Grazing on Giant Kelp Microscopic Phases and the Recruitment Success of Annual Populations of Macrocystis Pyrifera (Laminariales, Phaeophyta) in Southern Chile. J. Phycol 47, 252–258. doi: 10.1111/j.1529-8817.2010.00955.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hernández-Carmona G., Hughes B., Graham M. H. (2006). Reproductive Longevity of Drifting Kelp Macrocystis Pyrifera (Phaeophyceae) in Monterey Bay, USA. J. Phycol. 42, 1199–1207. doi: 10.1111/j.1529-8817.2006.00290.x

CrossRef Full Text | Google Scholar

Hernández-Carmona G., Robledo D., Serviere-Zaragoza E. (2001). Effect of Nutrient Availability on Macrocystis Pyrifera Recruitment Survival Near Its Southern Limit of Baja California. Botanica Marina 44, 221–229. doi: 10.1515/BOT.2001.029

CrossRef Full Text | Google Scholar

Hinojosa I. A., Pizarro M., Ramos M., Thiel M. (2010). Spatial and Temporal Distribution of Floating Kelp in the Channels and Fjords of Southern Chile. Estuar Coastal Shelf Sci. 87 (3), 367–377. doi: 10.1016/j.ecss.2009.12.010

CrossRef Full Text | Google Scholar

Hinton H. E. (1968). Reversible Suspension of Metabolism and the Origin of Life. Proc. R. Soc. Great Britain 171, 43–57. doi: 10.1098/rspb.1968.0055

CrossRef Full Text | Google Scholar

Hobday A. J. (2000). Abundance and Dispersal of Drifting Kelp Macrocystis Pyrifera Rafts in the Southern California Bight. Marine Ecol. Prog. Ser. 195, 101–116. doi: 10.3354/meps195101

CrossRef Full Text | Google Scholar

Hochachka P. W., Guppy M. (1987). Metabolic Arrest and the Control of Biological Time (Cambridge, Massachusetts: Harvard University Press).

Google Scholar

Hoffman A. J. (1987). The Arrival of Seaweed Propagules at the Shore: A Review. Botanica Marina 30, 151–165. doi: 10.1515/botm.1987.30.2.151

CrossRef Full Text | Google Scholar

Hoffman A. J., Avila M., Santelices B. (1984). Interactions of Nitrate and Phosphate on the Development of Microscopic Stages of Lessonia Nigrescens Bory (Phaeophyta). J. Exp. Marine Biol. Ecol. 78, 177–186. doi: 10.1016/0022-0981(84)90078-9

CrossRef Full Text | Google Scholar

Hoffman A. J., Camus P. (1989). Sinking Rates and Viability of Spores From Benthic Algae in Central Chile. J. Exp. Marine Biol. Ecol. 126, 281– 291. doi: 10.1016/0022-0981(89)90193-7

CrossRef Full Text | Google Scholar

Hoffman A. J., Santelices B. (1991). Banks of Microscopic Forms: Hypotheses on Their Functioning and Comparisons With Seed Banks. Marine Ecol. Prog. Ser. 79, 185–194. doi: 10.3354/meps079185

CrossRef Full Text | Google Scholar

Hollibaugh J. T., Seibert D. L. R., Thomas W. H. (1981). Observations on the Survival and Germination of Resting Spores of Three Chaetoceros (Bacillariophyceae) Species. J. Phycol. 171, 1–9. doi: 10.1111/j.1529-8817.1981.tb00812.x

CrossRef Full Text | Google Scholar

Hondolero D., Edwards M. S. (2017). Physical and Biological Characteristics of Kelp Forests in Kachemak Bay. Alaska. Marine Biol. 164, 81–93. doi: 10.1007/s00227-017-3111-3

CrossRef Full Text | Google Scholar

Hoos J. P. J. (2015). “Climate Change Impacts on the Kelp Life History Cycle,” in Ph.D. Thesis (Vancouver: University of British Columbia), 87 pp.

Google Scholar

Hsiao S. I. C., Druehl L. D. (1973). Environmental Control of Gametogenesis in Laminaria Saccharina. IV. In Situ Dev. gametophytes young sporophytes. J. Phycol 9, 160–164. doi: 10.1111/j.1529-8817.1973.tb04073.x

CrossRef Full Text | Google Scholar

Hubbard C. B., Garbary D. J., Kim K. Y., Chiasson D. M. (2004). Host Specificity and Growth of Kelp Gametophytes Symbiotic With Filamentous Red Algae (Ceramiales, Rhodophyta). Helgol. Mar. Res. 58 (1), 18–25. doi: 10.1007/s10152-003-0162-2

CrossRef Full Text | Google Scholar

IPCC (2013). Climate Change 2013: The Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge: Cambridge University Press).

Google Scholar

Izquierdo J., Pérez-Ruzafa I. M., Gallardo T. (2002). Effect of Temperature and Photon Fluence Rate on Gametophytes and Young Sporophytes of Laminaria Ochroleuca Pylaie. Helgoland Marine Res. 55, 285–292. doi: 10.1007/s10152-001-0087-6

CrossRef Full Text | Google Scholar

Jackson G. A. (1977). Nutrients and Production of Giant Kelp, Macrocystis Pyrifera, Off Southern California. Limnology Oceanogr. 22, 979–995. doi: 10.4319/lo.1977.22.6.0979

CrossRef Full Text | Google Scholar

Jackson G. A., Winant C. D. (1983). Effect of a Kelp Forest on Coastal Currents. Continental Shelf Res. 2, 75–80. doi: 10.1016/0278-4343(83)90023-7

CrossRef Full Text | Google Scholar

Jeon B. H., Yang K. M., Kim J. H. (2015). Changes in Macroalgal Assemblage With Sea Urchin Density on the East Coast of South Korea. Algae. 30, 139–146. doi: 10.4490/algae.2015.30.2.139

CrossRef Full Text | Google Scholar

Johnson C. R., Banks S. C., Barrett N. S., Cazassus F., Dunstan P. K., Edgar G. J., et al. (2011). Climate Change Cascades: Shifts in Oceanography, Species' Ranges and Subtidal Marine Community Dynamics in Eastern Tasmania. J. Exp. Marine Biol. Ecol. 400 (1-2), 17–32. doi: 10.1016/j.jembe.2011.02.032

CrossRef Full Text | Google Scholar

Justic D., Rabalais N. N., Turner R. E. (1997). Impacts of Climate Change on Net Productivity of Coastal Waters: Implications for Carbon Budgets and Hypoxia. Climate Res. 8 (3), 225–237. doi: 10.3354/cr008225

CrossRef Full Text | Google Scholar

Kain J. M. (1964). Aspects of the Biology of Laminaria Hyperborea III. Survival and Growth of Gametophytes. J. Marine Biol. Assoc. UK 44, 415–453. doi: 10.1017/S0025315400024929

CrossRef Full Text | Google Scholar

Keeley J. E. (1987). Role of Fire in Seed Germination of Woody Taxa in California Chaparral. Ecology 68, 434–443. doi: 10.2307/1939275

CrossRef Full Text | Google Scholar

Kidder K. A. (2006). “Ecology and Life History or Nereocystis Leutkeana in the South Slough Estuary,” in MS Thesis (Eugene: University of Oregon), 101 pp.

Google Scholar

Kim J. K., Kraemer G. P., Yarish C. (2015). Use of Sugar Kelp Aquaculture in Long Island Sound and the Bronx River Estuary for Nutrient Extraction. Marine Ecol. Prog. Ser. 531, 155–166. doi: 10.3354/meps11331

CrossRef Full Text | Google Scholar

Kinlan B. P., Graham M. H., Sala E., Dayton P. K. (2003). Arrested Development of Giant Kelp (Macrocystis Pyrifera, Phaeophyceae) Embryonic Sporophytes: A Mechanism for Delayed Recruitment in Perennial Kelps? J. Phycol. 39, 1–12. doi: 10.1046/j.1529-8817.2003.02087.x

CrossRef Full Text | Google Scholar

Klinger T. (1984). “Allocation of the Blade Surface Area to Meiospore Production in the Annual and Perennial Representatives of the Genus Laminaria,” in Master’s Thesis (Kelowna: University of British Columbia).

Google Scholar

Konar B. K., Edwards M. S., Efird T. (2015). Local Habitat and Regional Oceanographic Influence on Fish Distribution Patterns in the Diminishing Kelp Forests Across the Aleutian Archipelago. Environ. Biol. Fishes. 98, 1935–1951. doi: 10.1007/s10641-015-0412-6

CrossRef Full Text | Google Scholar

Konar B. K., Edwards M. S., Estes J. A. (2014). Biological Interactions Maintain the Boundaries Between Kelp Forests and Urchin Barrens in the Aleutian Archipelago. Hydrobiologia 724, 91–107. doi: 10.1007/s10750-013-1727-y

CrossRef Full Text | Google Scholar

Kopczak C. D., Zimmerman R. C., Kremer J. N. (1991). Variation in Nitrogen Physiology and Growth Among Geographically Isolated Populations of the Giant Kelp. Macrocystis Pyrifera (Phaeophyta). J. Phycol 27, 149–158. doi: 10.1111/j.0022-3646.1991.00149.x

CrossRef Full Text | Google Scholar

Krumhansl K. A., Okamoto D. K., Rassweiler A., Novak M., Bolton J. J., Cavanaugh K. C., et al. (2016). Global Patterns of Kelp Forest Change Over the Past Half-Century. Proc. Natl. Acad. Sci. 113, 13785 – 13790. doi: 10.1073/pnas.1606102113

CrossRef Full Text | Google Scholar

Ladah L. B., Zertuche-González J. A. (2007). Survival of Microscopic Stages of a Perennial Kelp (Macrocystis Pyrifera) From the Center and the Southern Extreme of Its Range in the Northern Hemisphere After Exposure to Simulated El Niño Stress. Marine Biol. 152 (3), 677–686. doi: 10.1007/s00227-007-0723-z

CrossRef Full Text | Google Scholar

Ladah L. B., Zertuche-González J. A., Hernández-Carmona G. (1999). Giant Kelp (Macrocystis Pyrifera, Phaeophyceae) Recruitment Near Its Southern Limit in Baja California After Mass Disappearance During ENSO 1997-1998. J. Phycol 35, 1106–1112. doi: 10.1046/j.1529-8817.1999.3561106.x

CrossRef Full Text | Google Scholar

Laeseke P., Bartsch I., Bischof K. (2019). Effects of Kelp Canopy on Underwater Light Climate and Viability of Brown Algal Spores in Kongsfjorden (Spitsbergen). Polar Biol. 42 (8), 1511–1527. doi: 10.1007/s00300-019-02537-w

CrossRef Full Text | Google Scholar

Leck M. A., Parker V. T., Simpson R. L. (1989). Ecology of Soil Seed Banks (San Diego, California, USA: Academic Press).

Google Scholar

Lemay M. A., Davis K. M., Martone P. T., Parfrey L. W. (2021). Kelp-Associated Microbiota Are Structured by Host Anatomy1. J. Phycol. 57 (4), 1119–1130. doi: 10.1111/jpy.13169

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemay M. A., Martone P. T., Keeling P. J., Burt J. M., Krumhansl K. A., Sanders R. D., et al. (2018). Sympatric Kelp Species Share a Large Portion of Their Surface Bacterial Communities. Environ. Microbiol. 20 (2), 658–670. doi: 10.1111/1462-2920.13993

PubMed Abstract | CrossRef Full Text | Google Scholar

Leonard G. H. (1994). Effwct of Bat Star Asterina Miniata (Brandt) on Recruitment of the Giant Kelp Macrocystis Pyrifera C. Agardh. J. Exp. Marine Biol. Ecol. 179, 81–98. doi: 10.1016/0022-0981(94)90018-3

CrossRef Full Text | Google Scholar

Levitus S., Antonov J., Boyer T. (2005). Warming of the World Ocean 1955-2003. Geophysical. Res. Lett. 32, 1–4. doi: 10.1029/2004GL021592

CrossRef Full Text | Google Scholar

Lewis W. M. J. (1985). Nutrient Scarcity as an Evolutionary Cause of Haploidy. Am. Nat. 125, 692– 701. doi: 10.1086/284372

CrossRef Full Text | Google Scholar

Lewis R. J., Green M. K., Afzal M. E. (2013). Effects of Chelated Iron on Oogenesis and Vegetative Growth of Kelp Gametophytes (P Haeophyceae). Phycol Res. 61 (1), 46–51. doi: 10.1111/j.1440-1835.2012.00667.x

CrossRef Full Text | Google Scholar

Liesner D., Fouqueau L., Valero M., Roleda M. Y., Pearson G. A., Bischof K., et al. (2020). Heat Stress Responses and Population Genetics of the Kelp Laminaria Digitata (Phaeophyceae) Across Latitudes Reveal Differentiation Among North Atlantic Populations. Ecol. Evol. 10 (17), 9144–9177. doi: 10.1002/ece3.6569

PubMed Abstract | CrossRef Full Text | Google Scholar

Lind A. C., Konar B. (2017). Effects of Abiotic Stressors on Kelp Early Life-History Stages. Algae 32 (3), 223–233. doi: 10.4490/algae.2017.32.8.7

CrossRef Full Text | Google Scholar

Lin J. D., Lemay M. A., Parfrey L. W. (2018). Diverse Bacteria Utilize Alginate Within the Microbiome of the Giant Kelp Macrocystis Pyrifera. Front. Microbiol. 9, 1914. doi: 10.3389/fmicb.2018.01914

PubMed Abstract | CrossRef Full Text | Google Scholar

Lubchenco J., Cubit J. (1980). Heteromorphic Life Histories of Certain Marine Algae as Adaptations to Variations in Herbivory. Ecology 64, 1116–1123. doi: 10.2307/1937822

CrossRef Full Text | Google Scholar

Lüning K. (1980). Critical Levels of Light and Temperature Regulating the Gametogenesis of Three Laminaria Species (Phaeophyceae). J. Phycol 16, 1–15. doi: 10.1111/j.1529-8817.1980.tb02992.x

CrossRef Full Text | Google Scholar

Lüning K., Neushul M. (1978). Light and Temperature Demands for Growth and Reproduction of Laminarian Gametophytes in Southern and Central California. Marine Biol. 45, 297–309. doi: 10.1007/BF00391816

CrossRef Full Text | Google Scholar

Lüthi D., Le Floc M., Bereite B., Blunier T., Barnola J. M., Siegenthaler U., et al. (2008). High-Resolution Carbon Dioxide Concentration Record 650,000–800,000 Years Before Present. Nature 453 (7193), 379–382. doi: 10.1038/nature06949

PubMed Abstract | CrossRef Full Text | Google Scholar

Macaya E. C., Boltaña S., Hinojosa I. A., Macchiavello J. E., Valdivia N. A., Vásquez N. R., et al. (2005). Presence of Sporophylls in Floating Kelp Rafts of Macrocystis Spp. (Phaeophyceae) Along the Chilean Pacific Coast. J. Phycol. 41, 913–922. doi: 10.1111/j.1529-8817.2005.00118.x

CrossRef Full Text | Google Scholar

Macaya E. C., Zuccarello G. C. (2010). Genetic Structure of the Giant Kelp Macrocystis Pyrifera Along the Southeastern Pacific. Marine Ecol. Prog. Ser. 420, 103–112. doi: 10.3354/meps08893

CrossRef Full Text | Google Scholar

Maier G. (1990). Spatial Distribution of Resting Stages, Rate of Emergence From Diapause, and Times to Adulthood and to the Appearance of the First Clutch in Three Species of Cyclopoid Copepods. Hydrobiologia 206, 11–18. doi: 10.1007/BF00018965

CrossRef Full Text | Google Scholar

Maier I., Hertweck C., Boland W. (2001). Stereochemical Specificity of Lamoxirene, the Sperm-Releasing Pheromone in Kelp (Laminariales, Phaeophyceae). Biol. Bull. 201 (2), 121–125. doi: 10.2307/1543327

PubMed Abstract | CrossRef Full Text | Google Scholar

Martins N., Pearson G. A., Gouveia L., Tavares A. I., Serrao E. A., Bartsch I. (2019). Hybrid Vigour for Thermal Tolerance in Hybrids Between the Allopatric Kelps Laminaria Digitata and L. Pallida (Laminariales, Phaeophyceae) With Contrasting Thermal Affinities. Eur. J. Phycol. 54 (4), 548–561. doi: 10.1080/09670262.2019.1613571

CrossRef Full Text | Google Scholar

Martins N., Tanttu H., Pearson G. A., Serrão E. A., Bartsch I. (2017). Interactions of Daylength, Temperature and Nutrients Affect Thresholds for Life Stage Transitions in the Kelp Laminaria Digitata (Phaeophyceae). Botanica Marina 60 (2), 109–121. doi: 10.1515/bot-2016-0094

CrossRef Full Text | Google Scholar

Matson P. G., Edwards M. S. (2006). Stipe Hollowing in Eisenia Arborea: Variation Across a Latitudinal Gradient. Phycologia 45, 343–348. doi: 10.2216/05-41.1

CrossRef Full Text | Google Scholar

Matson P. G., Edwards M. S. (2007). Effects of Ocean Temperature on the Southern Range Limits of Two Understory Kelps, Pterygophora Californica and Eisenia Arborea, at Multiple Life-Stages. Marine Biol. 151, 1941–1949. doi: 10.1007/s00227-007-0630-3

CrossRef Full Text | Google Scholar

McPherson M. L., Finger D. J., Houskeeper H. F., Bell T. W., Carr M. H., Rogers-Bennett L., et al. (2021). Large-Scale Shift in the Structure of a Kelp Forest Ecosystem Co-Occurs With an Epizootic and Marine Heatwave. Commun. Biol. 4 (1), 1–9. doi: 10.1038/s42003-021-01827-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Metzger J. R., Konar B., Edwards M. S. (2019). Assessing a Macroalgal Foundation Species: Community Variation With Shifting Algal Assemblages. Marine Biol. 166, 156. doi: 10.1007/s00227-019-3606-1

CrossRef Full Text | Google Scholar

Miller G. H., Alley R. B., Brigham-Grette J., Fitzpatrick J. J., Polyak L., Serreze M. C., et al. (2010). Arctic Amplification: Can the Past Constrain the Future? Quat Sci. Rev. 29, 1779–1790. doi: 10.1016/j.quascirev.2010.02.008

CrossRef Full Text | Google Scholar

Miller K. A., Estes J. A. (1989). Western Range Extension for Nereocystis Leutkeana in the North Pacific Ocean. Botanica Marina 32, 535–538. doi: 10.1515/botm.1989.32.6.535

CrossRef Full Text | Google Scholar

Miller R. J., Reed D. C., Brzezinski M. A. (2011). Partitioning of Primary Production Among Giant Kelp (Macrocystis Pyrifera), Understory Macroalgae, and Phytoplankton on a Temperate Reef. Limnol. Oceanogr. 56, 119–132. doi: 10.4319/lo.2011.56.1.0119

CrossRef Full Text | Google Scholar

Minich J. J., Morris M., Brown M., Doane M., Edwards M. S., Michael T. P., et al. (2018). Elevated Temperature Drives Kelp Microbiome Dysbiosis, While Elevated Carbon Dioxide Induces Water Microbiome Disruption. PloS One 13 (2), e0192772. doi: 10.1371/journal.pone.0192772

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohring M. B., Wernberg T., Wright J. T., Connell S. D., Russell B. D. (2014). Biogeographic Variation in Temperature Drives Performance of Kelp Gametophytes During Warming. Marine Ecol. Prog. Ser. 513, 85–96. doi: 10.3354/meps10916

CrossRef Full Text | Google Scholar

Molinos J. G., Burrows M. T., Poloczanska E. S. (2017). Ocean Currents Modify the Coupling Between Climate Change and Biogeographical Shifts. Sci. Rep. 7, 1332. doi: 10.1038/s41598-017-01309-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Morita T., Kurashima A., Maegawa M. (2003). Temperature Requirements for the Growth and Maturation of the Gametophytes of Undaria Pinnatifida and U. Undarioides (Laminariales, Phaeophyceae). Phycol Res. 51 (3), 154–160. doi: 10.1046/j.1440-1835.2003.t01-1-00305.x

CrossRef Full Text | Google Scholar

Morris M. M., Dinsdale E., Haggerty J. M., Edwards M. S. (2016). Nearshore Pelagic Microbial Community Abundance Affects Recruitment Success of Giant Kelp, Macrocystis Pyrifera. Front. Microbiol. 7, 1800. doi: 10.3389/fmicb.2016.01800

PubMed Abstract | CrossRef Full Text | Google Scholar

Morris R. L., Hale R., Strain E. M., Reeves S. E., Vergés A., Marzinelli E. M., et al. (2020). Key Principles for Managing Recovery of Kelp Forests Through Restoration. BioScience 70 (8), 688–698. doi: 10.1093/biosci/biaa058

CrossRef Full Text | Google Scholar

Müller R., Wiencke C., Bischof K. (2008). Interactive Effects of UV Radiation and Temperature on Microstages of Laminariales (Phaeophyceae) From the Arctic and North Sea. Climate Res. 37 (2-3), 203–213. doi: 10.3354/cr00762

CrossRef Full Text | Google Scholar

Müller R., Wiencke C., Bischof K., Krock B. (2009). Zoospores of Three Arctic Laminariales Under Different UV Radiation and Temperature Conditions: Exceptional Spectral Absorbance Properties and Lack of Phlorotannin Induction. Photochem. Photobiol. 85 (4), 970–977. doi: 10.1111/j.1751-1097.2008.00515.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Muñoz V., Hernández-González M. C., Buschmann A. H., Graham M. H., Vásquez J. A. (2004). Variability in Per Capita Oogonia and Sporophyte Production From Giant Kelp Gametophytes (Macrocystis Pyrifera, Phaeophyceae). Rev. Chil. Hist. Natural 77, 639–647.

Google Scholar

Muth A. F. (2012). Effects of Zoospore Aggregation and Substrate Rugosity on Kelp Recruitment Success. J. Phycol. 48 (6), 1374–1379. doi: 10.1111/j.1529-8817.2012.01211.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Muth A. F., Bonsell C., Dunton K. H. (2021). Inherent Tolerance of Extreme Seasonal Variability in Light and Salinity in an Arctic Endemic Kelp (Laminaria Solidungula). J. Phycol 57 (5), 1554–1562. doi: 10.1111/jpy.13187

PubMed Abstract | CrossRef Full Text | Google Scholar

Muth A. F., Graham M. H., Lane C. E., Harley C. D. G. (2019). Recruitment Tolerance to Increased Temperature Present Across Multiple Kelp Clades. Ecology 100 (3), e02594. doi: 10.1002/ecy.2594

PubMed Abstract | CrossRef Full Text | Google Scholar

Neal E. D., Hood E., Smikrud K. (2010). Contribution of Glacier Runoff to Freshwater Discharge Into the Gulf of Alaska. Geophys Res. Lett. 37 (6), 1–5. doi: 10.1029/2010GL042385

CrossRef Full Text | Google Scholar

Nelson W. A. (2005). Life History and Growth in Culture of the Endemic New Zealand Kelp Lessonia Variegata J. Agardh in Response to Differing Regimes of Temperature, Photoperiod and Light. J. Appl. Phycol. 17 (1), 23–28. doi: 10.1007/s10811-005-5521-8

CrossRef Full Text | Google Scholar

Neushul M. (1972). Underwater Microscopy With an Encased Incident-Light Dipping-Cone Microscope. J. Microsc. 95, 421–422. doi: 10.1111/j.1365-2818.1972.tb01043.x

CrossRef Full Text | Google Scholar

North W. (1994). “Review of Macrocystis Biology,” in Biology of Economic Algae. Ed. Akatsuka I. (Netherlands: Academic Publishing), pp 447–pp 527.

Google Scholar

Norton T. A. (1992). Dispersal by Macroalgae. Br. Phycol. J. 27, 293–301. doi: 10.1080/00071619200650271

CrossRef Full Text | Google Scholar

Novaczek I. (1984). Response of Gametophytes of Ecklonia Radiata (Laminariales) to Temperature in Saturating Light. Marine Biol. 82, 241–245. doi: 10.1007/BF00392405

CrossRef Full Text | Google Scholar

Olischläger M., Bartsch I., Gutow L., Wiencke C. (2012). Effects of Ocean Acidification on Different Life-Cycle Stages of the Kelp Laminaria Hyperborea (Phaeophyceae). Botanica Marina 55 (5), 511–525. doi: 10.1515/bot-2012-0163

CrossRef Full Text | Google Scholar

Oliver E. C. J., Benthuysen J. A., Bindoff N. L., Hobday A. J., Holbrook N. J., Mundy C. N., et al. (2017). The Unprecedented 2015/16 Tasman Sea Marine Heatwave. Nat. Commun. 8, 16101. doi: 10.1038/ncomms16101

PubMed Abstract | CrossRef Full Text | Google Scholar

O’Neel S., Hood E., Bidlack A. L., Fleming S. W., Arimitsu M. L., Arendt A., et al. (2015). Icefield-To-Ocean Linkages Across the Northern Pacific Coastal Temperate Rainforest Ecosystem. Bioscience 65, 499–512. doi: 10.1093/biosci/biv027

CrossRef Full Text | Google Scholar

Oppliger L. V., Correa J. A., Engelen A. H., Tellier F., Vieira V., Faugeron S., et al. (2012). Temperature Effects on Gametophyte Life-History Traits and Geographic Distribution of Two Cryptic Kelp Species. PloS One 7, e39289. doi: 10.1371/journal.pone.0039289

PubMed Abstract | CrossRef Full Text | Google Scholar

Paine E. R., Schmid M., Gaitán-Espitia J. D., Castle J., Jameson I., Sanderson J. C., et al. (2021). Narrow Range of Temperature and Irradiance Supports Optimal Development of Lessonia Corrugata (Ochrophyta) Gametophytes: Implications for Kelp Aquaculture and Responses to Climate Change. J. Appl. Phycol. 33 (3), 1721–1730. doi: 10.1007/s10811-021-02382-7

CrossRef Full Text | Google Scholar

Parnell P. E. (2015). The Effects of Seascape Pattern on Algal Patch Structure, Sea Urchin Barrens, and Ecological Processes. J. Exp. Marine Biol. Ecol. 465, 64–76. doi: 10.1016/j.jembe.2015.01.010

CrossRef Full Text | Google Scholar

Pearse J. S., Hines A. H. (1979). Expansion of a Central California Kelp Forest Following the Mass Mortality of Sea Urchins. Marine Biol. 51, 83–91. doi: 10.1007/BF00389034

CrossRef Full Text | Google Scholar

Perrot V., Richerd S., Valero M. (1991). Transition From Haploidy to Diploidy. Nature 351, 315–317. doi: 10.1038/351315a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Peters A. F., Couceiro L., Tsiamis K., Küpper F. C., Valero M. (2015). Barcoding of Cryptic Stages of Marine Brown Algae Isolated From Incubated Substratum Reveals High Diversity in Acinetosporaceae (Ectocarpales, Phaeophyceae). Cryptogamie Algologie 36 (1), 3–29. doi: 10.7872/crya.v36.iss1.2015.3

CrossRef Full Text | Google Scholar

Pfister C. A., Altabet M. A., Weigel B. L. (2019). Kelp Beds and Their Local Effects on Seawater Chemistry, Productivity, and Microbial Communities. Ecology 100, e02798. doi: 10.1002/ecy.2798

PubMed Abstract | CrossRef Full Text | Google Scholar

Phelps C. M., McMahon K., Bissett A., Bernasconi R., Steinberg P. D., Thomas T., et al. (2021). The Surface Bacterial Community of an Australian Kelp Shows Cross-Continental Variation and Relative Stability Within Regions. FEMS Microbiol. Ecol. 97 (7), p.fiab089. doi: 10.1093/femsec/fiab089

CrossRef Full Text | Google Scholar

Pierce S. M., Cowling R. M. (1991). Dynamics of Soil-Stored Seed Banks of Six Shrubs in Fire-Prone Dune Fynbos. J. Ecol. 79, 731–747. doi: 10.2307/2260664

CrossRef Full Text | Google Scholar

Pinter A. J., Lyman C. P., Willis J. S., Malan A., Wang L. C. H. (1984). Hibernation and Torpor in Mammals and Birds. J. Mammal 65, 172–175. doi: 10.2307/1381227

CrossRef Full Text | Google Scholar

Polovina J. J., Howell E. A., Abecassis M. (2008). Ocean's Least Productive Waters Are Expanding. Geophys Res. Lett. 35 (3). doi: 10.1029/2007GL031745

PubMed Abstract | CrossRef Full Text | Google Scholar

Provost E. J., Kelaher B. P., Dworjanyn S. A., Russell B. D., Connell S. D., Ghedini G., et al. (2017). Climate-Driven Disparities Among Ecological Interactions Threaten Kelp Forest Persistence. Global Change Biol. 23 (1), 353–361. doi: 10.1111/gcb.13414

CrossRef Full Text | Google Scholar

Raimondi P. T., Reed D. C., Gaylord B., Washburn L. (2004). Effects of Self-Fertilization in the Giant Kelp, Macrocystis Pyrifera. Ecol. 85, 3267–3276. doi: 10.1890/03-0559

CrossRef Full Text | Google Scholar

Ramirez-Puebla S. T., Weigel B. L., Jack L., Schlundt C., Pfister C. A., Welch J. L. M. (2020). Spatial Organization of the Kelp Microbiome at Micron Scales. bioRxiv 10 (1), 1–20. doi: 10.1101/2020.03.01.972083

CrossRef Full Text | Google Scholar

Ratcliff J. J., Soler-Vila A., Hanniffy D., Johnson M. P., Edwards M. D. (2017). Optimisation of Kelp (Laminaria Digitata) Gametophyte Growth and Gametogenesis: Effects of Photoperiod and Culture Media. J. Appl. Phycol. 29, 1957–1966. doi: 10.1007/s10811-017-1070-1

CrossRef Full Text | Google Scholar

Reed D. C. (1990). The Effects of Variable Settlement on Early Competition Patterns of Kelp Recruitment. Ecology 71, 776–787. doi: 10.2307/1940329

CrossRef Full Text | Google Scholar

Reed D. C., Amsler C. D., Ebeling A. W. (1992). Dispersal in Kelps: Factors Affecting Spore Swimming and Competency. Ecology 73, 1577–1585. doi: 10.2307/1940011

CrossRef Full Text | Google Scholar

Reed D. C., Anderson T. W., Ebeling A. W., Anghera M. (1997). The Role of Reproductive Synchrony in the Colonization Potential of Kelp. Ecology 78 (8), 2443–2457. doi: 10.1890/0012-9658(1997)078[2443:TRORSI]2.0.CO;2

CrossRef Full Text | Google Scholar

Reed D. C., Kinlan B. P., Raimondi P. T., Washburn L., Gaylord B., Drake P. T. (2006). “A Metapopulation Perspective on the Patch Dynamics of Giant Kelp in Southern California,” in Marine Metapopulations (Cambridge, Massachusetts: Academic Press), 353–386.

Google Scholar

Reed D. C., Laur D. R., Ebeling A. W. (1988). Variation in Algal Dispersal and Recruitment: The Importance of Episodic Events. Ecol. Monogr. 58, 321–335. doi: 10.2307/1942543

CrossRef Full Text | Google Scholar

Reed D. C., Neushul M., Ebeling A. W. (1991). Role of Settlement Density on Gametophyte Growth and Reproduction in the Kelps Pterygophora Californica and Macrocystis Pyrifera (Phaeophyceae). J. Phycol. 27, 361–366. doi: 10.1111/j.0022-3646.1991.00361.x

CrossRef Full Text | Google Scholar

Reed D. C., Raimondi P. T., Carr M. H., Goldwasser L. (2000). The Role of Dispersal and Disturbance in Determining Spatial Heterogeneity in Sedentary Organisms. Ecology 81, 2011–2026. doi: 10.1890/0012-9658(2000)081[2011:TRODAD]2.0.CO;2

CrossRef Full Text | Google Scholar

Reed D. C., Raimondi P. T., Washburn L., Gaylord B., Kinlan B. P., Drake P. T. (2004). “A Metapopulation Perspective on Patch Dynamics and Connectivity in Giant Kelp,” in Marine Metapopulations. Eds. Sale P., Kritzer J. (Cambridge, Massachusetts:Academic Press).

Google Scholar

Reed D. C., Schroeter S. C. (2004). Spore Supply and Habitat Availability as Sources of Recruitment Limitation in the Giant Kelp Macrocystis Pyrifera (Phaeophyceae). J. Phycol. 40, 275–284. doi: 10.1046/j.1529-8817.2004.03119.x

CrossRef Full Text | Google Scholar

Reed D. C., Schroeter S. C., Raimondi P. T. (2004). Spore Supply and Habitat Availability as Sources of Recruitment Limitation in Giant Help, Macrocystis Pyrifera. J. Phycol. 40, 275–284. doi: 10.1046/j.1529-8817.2004.03119.x

CrossRef Full Text | Google Scholar

Reed D., Washburn L., Rassweiler A., Miller R., Bell T., Harrer S. (2016). Extreme Warming Challenges Sentinel Status of Kelp Forests as Indicators of Climate Change. Nat. Commun. 7 (1), 1–7. doi: 10.1038/ncomms13757

CrossRef Full Text | Google Scholar

Reisdorph S. C., Mathis J. T. (2014). The Dynamic Controls on Carbonate Mineral Saturation States and Ocean Acidification in a Glacially Dominated Estuary. Estuar Coastal Shelf Sci. 144, 8–18. doi: 10.1016/j.ecss.2014.03.018

CrossRef Full Text | Google Scholar

Richardson A. J. (2008). In Hot Water: Zooplankton and Climate Change. ICES J. Marine Sci. 65 (3), 279–295. doi: 10.1093/icesjms/fsn028

CrossRef Full Text | Google Scholar

Roberson L. M., Coyer J. A. (2004). Variation in Blade Morphology of the Kelp Eisenia Arborea: Incipient Speciation Due to Local Water Motion? Marine Ecol. Prog. Ser. 282, 115–128. doi: 10.3354/meps282115

CrossRef Full Text | Google Scholar

Rodgers-Bennet L., Catton C. A. (2019). Marine Heat Wave and Multiple Stressors Tip Bull Kelp Forests to Sea Urchin Barren Grounds. Sci. Rep. 9.1, 1–9. doi: 10.1038/s41598-019-51114-y

CrossRef Full Text | Google Scholar

Roleda M. Y. (2009). Photosynthetic Response of Arctic Kelp Zoospores Exposed to Radiation and Thermal Stress. Photochem Photobiol Sci. 8 (9), 1302–1312. doi: 10.1039/b901098j

PubMed Abstract | CrossRef Full Text | Google Scholar

Roleda M. Y., Morris J. N., McGraw C. M., Hurd C. L. (2012). Ocean Acidification and Seaweed Reproduction: Increased CO 2 Ameliorates the Negative Effect of Lowered pH on Meiospore Germination in the Giant Kelp Macrocystis Pyrifera (Laminariales, Phaeophyceae). Global Change Biol. 18 (3), 854–864. doi: 10.1111/j.1365-2486.2011.02594.x

CrossRef Full Text | Google Scholar

Roleda M. Y., Wiencke C., Hanelt D., van de Poll W. H., Gruber A. (2005). Sensitivity of Laminariales Zoospores From Helgoland (North Sea) to Ultraviolet and Photosynthetically Active Radiation: Implications for Depth Distribution and Seasonal Reproduction. Plant Cell Environ. 28 (4), 466–479. doi: 10.1111/j.1365-3040.2005.01288.x

CrossRef Full Text | Google Scholar

Rothäusler E., Gómez I., Hinojosa I. A., Karsten U., Tala F., Thiel M. (2009). Effect of Temperature and Grazing on Growth and Reproduction of Floating Macrocystis Spp. (Phaeophyceae) Along a Latitudinal Gradient. J. Phycol. 45, 547–559. doi: 10.1111/j.1529-8817.2009.00676.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Rubechon M., Couceiro L., Peters A. F., Destombe C., Valero M. (2014). Examining the Bank of Microscopic Stages in Kelps Using Culturing and Barcoding. Eur. J. Phycol. 49, 128–133. doi: 10.1080/09670262.2014.892635

CrossRef Full Text | Google Scholar

Sabine C. L., Feely R. A., Gruber N., Key R. M., Lee K., Bullister J. L., et al. (2004). The Oceanic Sink for Anthropogenic CO2. Sci. 305, 367–371. doi: 10.1126/science.1097403

CrossRef Full Text | Google Scholar

Scavia D., Field J. C., Boesch D. F., Buddemeier R. W., Burkett V., Cayan D. R., et al. (2002). Climate Change Impacts on US Coastal and Marine Ecosystems. Estuaries 25 (2), 149–164. doi: 10.1007/BF02691304

CrossRef Full Text | Google Scholar

Scheffer M., Brovkin V., Cox P. M. (2006). Positive Feedback Between Global Warming and Atmospheric CO2 Concentration Inferred From Past Climate Change. Geophys Res. Lett. 33, L10702. doi: 10.1029/2005GL025044

CrossRef Full Text | Google Scholar

Scheibling R. E., Hennigar A. W., Balch T. (1999). Destructive Grazing, Epiphytism, and Disease: The Dynamics of Sea Urchin - Kelp Interactions in Nova Scotia. Can. J. Fish Aquat. Sci. 56, 2300–2314. doi: 10.1139/f99-163

CrossRef Full Text | Google Scholar

Schiel D. R., Foster M. S. (2015). The Biology and Ecology of Giant Kelp. (Oakland: University of California Press), 416 pp.

Google Scholar

Schoenrock K. M., McHugh T. A., Krueger-Hadfield S. A. (2021). Revisiting the ‘Bank of Microscopic Forms’ in Macroalgal-Dominated Ecosystems. J. pf Phycol. 57, 14–29. doi: 10.1111/jpy.13092

CrossRef Full Text | Google Scholar

Seymour R. J., Tegner M. J., Dayton P. K., Parnell P. E. (1989). Storm Wave Induced Mortality of Giant Kelp, Macrocystis Pyrifera, in Southern California. Estuar Coastal Shelf Sci. 28 (3), 277–292. doi: 10.1016/0272-7714(89)90018-8

CrossRef Full Text | Google Scholar

Shukla P., Edwards M. S. (2017). Elevated Pco2 Is Less Detrimental That Increased Temperature to Early Development of the Giant Kelp, Macrocystis Pyrfera (Phaeophyceae, Lamnariales). Phycologia 56, 638–648. doi: 10.2216/16-120.1

CrossRef Full Text | Google Scholar

Silva P. C. (1992). Geographic Patterns of Diversity in Benthic Marine Algae. Pacific Sci. 46, 429–437.

Google Scholar

Silva C. F., Pearson G. A., Serrao E. A., Bartsch I., Martins N. (2022). Microscopic Life Stages of Arctic Kelp Differ in Their Resilience and Reproductive Output in Response to Arctic Seasonality. Eur. J. Phycol. pp, 1–15. doi: 10.1080/09670262.2021.2014983

CrossRef Full Text | Google Scholar

Slocum C. J. (1980). Differential Susceptibility to Grazers in Two Phases of an Intertidal Alga: Advantages of Heteromorphic Generations. J. Exp. Marine Biol. Ecol. 46, 99–110. doi: 10.1016/0022-0981(80)90095-7

CrossRef Full Text | Google Scholar

Smale D. A. (2020). Impacts of Ocean Warming on Kelp Forest Ecosystems. N. Phytol. 225, 1447–1454. doi: 10.1111/nph.16107

CrossRef Full Text | Google Scholar

Smale D. A., Wernberg T., Yunnie A. L., Vance T. (2015). The Rise of Laminaria Ochroleuca in the Western English Channel (UK) and Comparisons With Its Competitor and Assemblage Dominant Laminaria Hyperborea. Marine Ecol. 36 (4), 1033–1044. doi: 10.1111/maec.12199

CrossRef Full Text | Google Scholar

Small S., Edwards M. S. (2021). Thermal Tolerance may Slow, But Not Prevent, the Spread of Sargassum Horneri (Phaeophyceae) Along the California, USA and Baja California, MEX Coastline. J. Phycol. 57 (3), 903–915. doi: 10.1111/jpy.13148

PubMed Abstract | CrossRef Full Text | Google Scholar

Spalding H., Foster M. S., Heine J. N. (2003). Composition, Distribution, and Abundance of Deep-Water (>30 M) Macroalgae in Central California. J. Phycol. 39, 273–284. doi: 10.1046/j.1529-8817.2003.02010.x

CrossRef Full Text | Google Scholar

Spector M., Edwards M. S. (2020). Modelling the Impacts of Kelp Deforestation on Benthic Primary Production on Temperate Rocky Reefs. Algae 35 (3), 237–252. doi: 10.4490/algae.2020.35.8.19

CrossRef Full Text | Google Scholar

Spindel N. B., Lee L. C., Okmoto D. K. (2021). Zombies of the Nearshore: Metabolic Depression in Sea Urchin Barrens Associated With Food Deprivation. bioRixivb 102 (4), e01926. doi: 10.1002/bes2.1926

CrossRef Full Text | Google Scholar

Steneck R. S., Graham M. H., Bourque B. J., Corbett D., Erlandson J. M., Estes J. A., et al. (2002). Kelp Forest Ecosystems: Biodiversity, Stability, Resilience and Future. Environ. Conserv. 29, 436–459. doi: 10.1017/S0376892902000322

CrossRef Full Text | Google Scholar

Stevens C. L., Hurd C. L., Isachsen P. E. (2003). Modelling of Diffusion Boundary Layers in Subtidal Macroalgal Canopies: The Response to Waves and Currents. Aquat. Sci. 65, 81–91. doi: 10.1007/s000270300007

CrossRef Full Text | Google Scholar

Sullaway G., Edwards M. S. (2020). Impacts of the Non-Native Alga, Sargassum Horneri, on Benthic Primary Production in a California Kelp Forest. Marine Ecol. Prog. Series. 637, 45–57. doi: 10.3354/meps13231

CrossRef Full Text | Google Scholar

Sun C., Feng M., Matear R. J., Chamberlain M. A., Craig P., Ridgway K. R., et al. (2012). Marine Downscaling of a Future Climate Scenario for Australian Boundary Currents. J. Climate 25, 2947–2962. doi: 10.1175/JCLI-D-11-00159.1

CrossRef Full Text | Google Scholar

Swanson A. K., Druehl L. D. (2000). Differential Meiospore Size and Tolerance of Ultraviolet Light Stress Within and Among Kelp Species Along a Depth Gradient. Marine Biol. 136, 657–664. doi: 10.1007/s002270050725

CrossRef Full Text | Google Scholar

Tøien Ø., Blake J., Edgar D. M., Grahn D. A., Hdller H. C., Barnes B. M. (2011). Hibernation in Black Bears: Independence of Metabolic Suppression From Body Temperature. Science 331, 906–909. doi: 10.1126/science.1199435

PubMed Abstract | CrossRef Full Text | Google Scholar

Tauber M. J., Tauber C. A. (1978). “Evolution of Phenological Strategies in Insects: A Comparative Approach With Eco-Physiological and Genetic Considerations,” in Evolution of Insect Migration and Diapause. Ed. Dingle H. (New York, NY: Springer).

Google Scholar

Teagle H., Hawkins S. J., Moore P. J., Smale D. A. (2017). The Role of Kelp Species as Biogenic Habitat Formers in Coastal Marine Ecosystems. J. Exp. Marine Biol. Ecol. 492, 81–98. doi: 10.1016/j.jembe.2017.01.017

CrossRef Full Text | Google Scholar

Tegner M. J., Dayton P. K. (1987). El Niño Effects on Southern California Kelp Forest Communities. Adv. Ecol. Res. 17, 243–279. doi: 10.1016/S0065-2504(08)60247-0

CrossRef Full Text | Google Scholar

Tegner M. J., Dayton P. K., Edwards P. B., Riser K. L., Chadwick D. B., Dean T. A., et al. (1995). Effects of a Large Sewage Spill on a Kelp Forest Community—Catastrophe or Disturbance. Marine Environ. Res. 40, 181–224. doi: 10.1016/0141-1136(94)00008-D

CrossRef Full Text | Google Scholar

Thiel M. (2003). “Rafting of Benthic Macrofauna: Important Factors Determining the Temporal Succession of the Assemblage on Detached Macroalgae” in Migrations and Dispersal of Marine Organisms (New York: Springer, Dordrecht), 49–57.

Google Scholar

Thornber C. S., Gaines S. D. (2004). Population Demographics in Species With Biphasic Life Cycles. Ecology 85 (6), 1661–1674. doi: 10.1890/02-4101

CrossRef Full Text | Google Scholar

tom Dieck I. T. (1993). Temperature Tolerance and Survival in Darkness of Kelp Gametophytes (Laminariales, Phaeophyta): Ecological and Biogeographical Implications. Marine Ecol. Prog. Ser. 100, 253–264. doi: 10.3354/meps100253

CrossRef Full Text | Google Scholar

Traiger S. B., Konar B. (2017). Supply and Survival: Glacial Melt Imposes Limitations at the Kelp Microscopic Life Stages. Botanical Marina 60, 603–617. doi: 10.1515/bot-2017-0039

CrossRef Full Text | Google Scholar

Tseng C. K., Ren K. Z., Wu C. Y. (1959). On the Discharge of Eggs and Spermatozoids of Laminaria Japonica and the Morphology of the Spermatozoids. Kexue Tongbao 4, 129–130.

Google Scholar

Turpen S., Hunt J. W., Anderson B. S., Pearse J. S. (1994). Population Structure, Growth, and Fecundity of the Kelp Forest Mysid Holmesimysis Costata in Monterey Bay. California J. Crustacean Biol. 14, 657–664. doi: 10.2307/1548859

CrossRef Full Text | Google Scholar

Underweood A. J., Fairweather P. G. (1989). Supply-Side Ecology and Benthic Marine Assemblages. Trends Ecol. Evol. 4, 16–20. doi: 10.1016/0169-5347(89)90008-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Vadas R. L. (1972). Ecological Implications of Culture Studies on Nereocystis Leutkeana. J. Phycol 8, 196–203. doi: 10.1111/j.1529-8817.1972.tb04025.x

CrossRef Full Text | Google Scholar

Vadas R. L., Steneck R. S. (1988). Zonation of Deep Water Benthic Algae in the Gulf of Maine. J. Phycol. 24, 338–346. doi: 10.1111/j.1529-8817.1988.tb04476.x

CrossRef Full Text | Google Scholar

Van den Hoek C. (1982). Phytogeographic Distribution Groups of Benthic Marine Algae in the North Atlantic Ocean. A Review of Experimental Evidence From Life History Studies. Helgoländer Meeresuntersuchungen 35 (2), 153–214. doi: 10.1007/BF01997551

CrossRef Full Text | Google Scholar

Van den Hoek C. (1987). The Possible Significance of Long-Range Dispersal for the Biogeography of Seaweeds. Helgoländer Meeresuntersuchungen 41 (3), 261–272. doi: 10.1007/BF02366191

CrossRef Full Text | Google Scholar

Vanderklift M. A., Wernberg T. (2008). Detached Kelps From Distant Sources Are a Food Subsidy for Sea Urchins. Oecologia 157, 327–335. doi: 10.1007/s00442-008-1061-7

PubMed Abstract | CrossRef Full Text | Google Scholar

VanMeter K., Edwards M. S. (2013). The Effects of Grazing on Kelp Zoospore Dispersal Potential. J. Phycol 49, 896–901. doi: 10.1111/jpy.12100

PubMed Abstract | CrossRef Full Text | Google Scholar

Veenhof R. J., Champion C., Dworjanyn S. A., Wernberg T., Minne A. J. P., Layton C., et al. Kelp Gametophytes in Changing Oceans. Oceanogr. Marine Biol. Annu. Rev. 60.

Google Scholar

Venable D. L., Lawlor L. (1980). Delayed Germination and Dispersal in Desert Annuals: Escape in Space and Time. Oeocologia 46, 272–282. doi: 10.1007/BF00540137

CrossRef Full Text | Google Scholar

Voerman S. E., Llera E., Rico J. M. (2013). Climate Driven Changes in Subtidal Kelp Forest Communities in NW Spain. Marine Environ. Res. 90, 119–127. doi: 10.1016/j.marenvres.2013.06.006

CrossRef Full Text | Google Scholar

Voosen P. (2020). Climate Change Spurs Global Speedup of Ocean Currents. Science 367, 612–613. doi: 10.1126/science.367.6478.612

PubMed Abstract | CrossRef Full Text | Google Scholar

Watanabe Y., Nishihara G. N., Tokunaga S., Terada R. (2014). The Effect of Irradiance and Temperature Responses and the Phenology of a Native Alga, Undaria Pinnatifida (Laminariales), at the Southern Limit of its Natural Distribution in Japan. J. Appl. Phycol. 26 (6), 2405–2415. doi: 10.1007/s10811-014-0264-z

CrossRef Full Text | Google Scholar

Weigel B. L., Pfister C. A. (2019). Successional Dynamics and Seascape-Level Patterns of Microbial Communities on the Canopy-Forming Kelps Nereocystis Luetkeana and Macrocystis Pyrifera. Front. Microbiol. 10, 346. doi: 10.3389/fmicb.2019.00346

PubMed Abstract | CrossRef Full Text | Google Scholar

Wernberg T., Bennett S., Babcock R. C., de Bettignies T., Cure K., Depczynski M., et al. (2016). Climate-Driven Regime Shift of a Temperate Marine Ecosystem. Science 353, 169–172. doi: 10.1126/science.aad8745

PubMed Abstract | CrossRef Full Text | Google Scholar

Wernberg T., Coleman M. A., Babcock R. C., Bell S. Y., Bolton J. J., Connell S. D., et al. (2019b). Biology and Ecology of the Globally Significant Kelp Ecklonia Radiata. Oceanogr. Marine Biol. doi: 10.1201/9780429026379-6

CrossRef Full Text | Google Scholar

Wernberg T., Krumhansl K., Filbee-Dexter K., Pedersen M. F. (2019a). “Status and Trends for the World’s Kelp Forests,” in World Seas: An Environmental Evaluation (Cambridge, Massachusetts: Academic Press), (pp. 57–78).

Google Scholar

Wernberg T., Thomsen M. S., Tuya F., Kendrick G. A., Staehr P. A., Toohey B. D. (2010). Decreasing Resilience of Kelp Beds Along a Latitudinal Temperature Gradient: Potential Implications for a Warmer Future. Ecol. Lett. 13, 685–694. doi: 10.1111/j.1461-0248.2010.01466.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wiencke C., Gómez I., Pakker H., Flores-Moya A., Altamirano M., Hanelt D., et al. (2000). Impact of UV-Radiation on Viability, Photosynthetic Characteristics and DNA of Brown Algal Zoospores: Implications for Depth Zonation. Marine Ecol. Prog. Ser. 197, 217–229. doi: 10.3354/meps197217

CrossRef Full Text | Google Scholar

Wilmers C. C., Estes J. A., Edwards M., Laidre K. L., Konar B. (2012). Do Trophic Cascades Affect the Storage and Flux of Atmospheric Carbon? An Analysis of Sea Otters and Kelp Forests. Front. Ecol. Environ. 10 (8), 409–415. doi: 10.1890/110176

CrossRef Full Text | Google Scholar

Wollschläger J., Neale P. J., North R. L., Striebel M., Zielinski O. (2021). Climate Change and Light in Aquatic Ecosystems: Variability & Ecological Consequences. Front. Marine Sci. 8, 506. doi: 10.3389/fmars.2021.688712

CrossRef Full Text | Google Scholar

Yoneshigue-Valentin Y. (1990). The Life Cycle of Laminaria Abyssalis (Laminariales, Phaeophyta) in Culture. Hydrobiologia 204/205, 461–466. doi: 10.1007/BF00040271

CrossRef Full Text | Google Scholar

Zacher K., Bernard M., Bartsch I., Wiencke C. (2016). Survival of Early Life History Stages of Arctic Kelps (Kongsfjorden, Svalbard) Under Multifactorial Global Change Scenarios. Polar Biol. 39 (11), 2009–2020. doi: 10.1007/s00300-016-1906-1

CrossRef Full Text | Google Scholar

Zacher K., Bernard M., Daniel Moreno A., Bartsch I. (2019). Temperature Mediates the Outcome of Species Interactions in Early Life-History Stages of Two Sympatric Kelp Species. Marine Biol. 166 (12), 1–16. doi: 10.1007/s00227-019-3600-7

CrossRef Full Text | Google Scholar

Zarco-Perello S., Bosch N. E., Bennett S., Vanderklift M. A., Wernberg T. (2021). Persistence of Tropical Herbivores in Temperate Reefs Constrains Kelp Resilience to Cryptic Habitats. J. Ecol. 109 (5), 2081–2094. doi: 10.1111/1365-2745.13621

CrossRef Full Text | Google Scholar

Žuljević A., Peters A. F., Nikolić V., Antolić B., Despalatović M., Cvitković I., et al. (2016). The Mediterranean Deep-Water Kelp Laminaria Rodriguezii Is an Endangered Species in the Adriatic Sea. Marine Biol. 163 (4), 1–12. doi: 10.1007/s00227-016-2821-2

CrossRef Full Text | Google Scholar

Keywords: climate change, gametophyte, kelp, microscopic, recruitment, sporophyte, zoospore

Citation: Edwards MS (2022) It’s the Little Things: The Role of Microscopic Life Stages in Maintaining Kelp Populations. Front. Mar. Sci. 9:871204. doi: 10.3389/fmars.2022.871204

Received: 07 February 2022; Accepted: 29 March 2022;
Published: 29 April 2022.

Edited by:

Michael Yu Roleda, University of the Philippines Diliman, Philippines

Reviewed by:

Stein Fredriksen, University of Oslo, Norway
Catherine Ann Pfister, The University of Chicago, United States
Inka Bartsch, Alfred Wegener Institute Helmholtz Centre for Polar and Marine Research (AWI), Germany

Copyright © 2022 Edwards. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Matthew S. Edwards, medwards@sdsu.edu

Download