harmful algal blooms and eutrophication nutrient sources, composition, and consequences

Upload: jca

Post on 06-Apr-2018

224 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    1/23

    704 2002 Estuarine Research Federation

    Estuaries Vol. 25, No. 4b, p. 704726 August 2002

    Harmful Algal Blooms and Eutrophication: Nutrient Sources,

    Composition, and Consequences

    DONALD M. ANDERSON1*, PATRICIA M. GLIBERT2, and JOANN M. BURKHOLDER3

    1 Biology Department, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts 025432 University of Maryland Center for Environmental Science, Horn Point Laboratory, P. O. Box

    775, Cambridge, Maryland 216133 Center for Applied Aquatic Ecology, North Carolina State University, 620 Hutton Street, Suite

    104, Raleigh, North Carolina 27606

    ABSTRACT: Although algal blooms, including those considered toxic or harmful, can be natural phenomena, the nature

    of the global problem of harmful algal blooms (HABs) has expanded both in extent and its public perception over the

    last several decades. Of concern, especially for resource managers, is the potential relationship between HABs and the

    accelerated eutrophication of coastal waters from human activities. We address current insights into the relationships

    between HABs and eutrophication, focusing on sources of nutrients, known effects of nutrient loading and reduction,

    new understanding of pathways of nutrient acquisition among HAB species, and relationships between nutrients andtoxic algae. Through specific, regional, and global examples of these various relationships, we offer both an assessment

    of the state of understanding, and the uncertainties that require future research efforts. The sources of nutrients poten-

    tially stimulating algal blooms include sewage, atmospheric deposition, groundwater flow, as well as agricultural and

    aquaculture runoff and discharge. On a global basis, strong correlations have been demonstrated between total phos-

    phorus inputs and phytoplankton production in freshwaters, and between total nitrogen input and phytoplankton pro-

    duction in estuarine and marine waters. There are also numerous examples in geographic regions ranging from the

    largest and second largest U.S. mainland estuaries (Chesapeake Bay and the Albemarle-Pamlico Estuarine System), to

    the Inland Sea of Japan, the Black Sea, and Chinese coastal waters, where increases in nutrient loading have been linked

    with the development of large biomass blooms, leading to anoxia and even toxic or harmful impacts on fisheries re-

    sources, ecosystems, and human health or recreation. Many of these regions have witnessed reductions in phytoplankton

    biomass (as chlorophyll a) or HAB incidence when nutrient controls were put in place. Shifts in species composition

    have often been attributed to changes in nutrient supply ratios, primarily N:P or N:Si. Recently this concept has been

    extended to include organic forms of nutrients, and an elevation in the ratio of dissolved organic carbon to dissolved

    organic nitrogen (DOC:DON) has been observed during several recent blooms. The physiological strategies by which

    different groups of species acquire their nutrients have become better understood, and alternate modes of nutrition

    such as heterotrophy and mixotrophy are now recognized as common among HAB species. Despite our increased un-

    derstanding of the pathways by which nutrients are delivered to ecosystems and the pathways by which they are assimilated

    differentially by different groups of species, the relationships between nutrient delivery and the development of blooms

    and their potential toxicity or harmfulness remain poorly understood. Many factors such as algal species presence/

    abundance, degree of flushing or water exchange, weather conditions, and presence and abundance of grazers contribute

    to the success of a given species at a given point in time. Similar nutrient loads do not have the same impact in different

    environments or in the same environment at different points in time. Eutrophication is one of several mechanisms by

    which harmful algae appear to be increasing in extent and duration in many locations. Although important, it is not the

    only explanation for blooms or toxic outbreaks. Nutrient enrichment has been strongly linked to stimulation of some

    harmful species, but for others it has not been an apparent contributing factor. The overall effect of nutrient over-

    enrichment on harmful algal species is clearly species specific.

    Introduction

    Algal blooms, including toxic events, can be nat-ural phenomena. Historically, indigenous tribesavoided shellfish at certain places or times of year(e.g., Lescarbot 1609 cited in Prakash et al. 1971),and the logs of early mariners such as Captains

    James Cook and George Vancouver (Vancouver

    * Corresponding author; fax: 508/457-2027; e-mail: [email protected].

    and Robinson 1798 cited in Prakash et al. 1971)

    describe discolored water and poisonous shellfish.Over the last several decades coastal regionsthroughout the world have experienced what ap-pears to be an escalation in the incidence ofblooms that are toxic or otherwise harmful. Com-monly called red tides, these events are nowgrouped under the descriptor harmful algalblooms or HABs. Although most of the species in-

    volved are plant-like, photosynthetic algae, a feware actually animal-like protozoans without the

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    2/23

    HABs and Eutrophication 705

    ability to photosynthesize on their own. HABs haveone unique feature in commonthey cause harm,either due to their production of toxins or to themanner in which the cells physical structure oraccumulated biomass affect co-occurring organ-isms and alter food web dynamics. Impacts of these

    phenomena include mass mortalities of wild andfarmed fish and shellfish; human illness and deathfrom toxic seafood or from toxin exposurethrough inhalation or water contact; illness anddeath of marine mammals, seabirds, and other an-imals; and alteration of marine habitats and tro-phic structure.

    A distinction must be made between two differ-ent types of HABsthose that involve toxins orharmful metabolites, such as toxins linked to wild-life death or human seafood poisonings, and those

    which are nontoxic but cause harm in other ways.Some algal toxins are extremely potent, and low-density blooms can be dangerous, sometimes caus-

    ing poisonings at concentrations as low as a fewhundred cells l1. Many HAB species that do notproduce toxins are able to cause harm through thedevelopment of high biomass, leading to foams orscums, the depletion of oxygen as blooms decay,or the destruction of habitat for fish or shellfish byshading of submerged vegetation.

    Eutrophication is the natural aging process ofaquatic ecosystems. The term was formerly usedmostly in reference to the natural aging of lakes

    wherein a large, deep, nutrient-poor lake eventu-ally becomes more nutrient-rich, more productive

    with plant and animal life, and slowly fills in tobecome a pond, then a marsh (Wetzel 1983). More

    recently, the term has been used to refer to cul-tural or accelerated eutrophication of lakes, rivers,estuaries, and marine waters, wherein the naturaleutrophication process is advanced by hundreds orthousands of years by human activities that add nu-trients (Burkholder 2000). Nixon (1995, p. 95) de-fined eutrophication as the process of increasedorganic enrichment of an ecosystem, generallythrough increased nutrient inputs.

    Two nutrients in human-derived sources, phos-phorus (P) and nitrogen (N), are of most concernin eutrophication. In freshwaters, P is the leastabundant among the nutrients needed in largequantity (macronutrients) by photosynthetic or-ganisms, so it is the primary nutrient that limitstheir growth (Schindler 1977). P can also limit orco-limit algal growth in estuarine and marine en-

    vironments that are sustaining high N inputs (Ru-dek et al. 1991; Fisher et al. 1992). In many tem-perate and polar coastal marine waters, N is themost important nutrient that limits primary pro-duction of photosynthetic organisms (Dugdale andGoering 1967; Glibert 1988). N is often the nutri-

    ent that first limits primary production at the es-tuarine interface between marine and freshwaterhabitats. In lower estuaries both N and P can co-limit phytoplankton production (Rudek et al.1991; Fisher et al. 1992). If improved sewage treat-ment reduces P loading within freshwater seg-

    ments of a given river system, corresponding re-ductions in freshwater phytoplankton blooms willallow more inorganic N to be transported down toestuarine segments where it can support largerblooms (Fisher et al. 1992; Mallin et al. 1993).Both N and P are considered here, and these nu-trients should be co-managed in the developmentof strategies to minimize HABs. Other nutrientssuch as silicon (Si) and iron (Fe) also can signifi-cantly influence the outcome of species domi-nance and the structure and abundance of phyto-plankton communities under cultural eutrophica-tion (Heckey and Kilham 1988; Wilhelm 1995).

    For more than 50 years scientists have recog-

    nized that noxious blooms of toxic or otherwiseharmful cyanobacteria (blue-green algae), themost common harmful algae in freshwater lakes,reservoirs, and slow flowing rivers, are stimulatedby P enrichment (reviewed in Schindler 1977;Smith 1983). These organisms can form rotting hy-perscum mats up to ca. 1 m thick, with billions ofcells ml1 and chlorophyll a (chl a; index of algalbiomass) as high as 3,000 g l1 (Zohary and Rob-erts 1989). Many species produce bioactive com-pounds, including potent hepatotoxins and neu-rotoxins that have caused livestock and wildlifedeath in most countries throughout the world(Skulberg et al. 1993; Codd et al. 1997) and, more

    rarely, death of humans as well (Chorus and Bar-tram 1999). The relationship between cyanobac-teria and P is sufficiently strong that in many lakesof moderate depth ( 10 m) with low abiotic tur-bidity, the spring-season concentration of total P inlakes (specifically, during lake overturn or total wa-ter column mixing) has been used with reasonablesuccess to predict the late summer maximum incyanobacterial biomass (as water-column chl a;

    Wetzel 1983). This relationship has also held inestuarine and brackish coastal waters of Scandina-

    via and Australia, where blooms of the toxic cya-nobacterium, Nodularia spumigena, have been re-lated to excessive P enrichment (Chorus and Bar-tram 1999).

    In freshwater reservoirs and rivers, mixing andflushing dynamics are more complex, and abioticturbidity from episodic sediment loading is appre-ciable. Light can be the primary resource limitingalgal growth, rather than nutrients. The increasedflow and mixing maintains relatively high nutrientsupplies, and P has not been used successfully topredict the occurrence and extent of late summer

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    3/23

    706 D. M. Anderson et al.

    cyanobacterial blooms (Canfield and Bachmann1981; Thornton et al. 1990). Modest success in un-derstanding nutrient stimulation of harmful algae,and in being able to reliably predict HABs fromnutrient inputs, has been achieved to date only forcyanobacteria in clear-water lakes of moderate

    depth and dependable mixing regimes. Reliableprediction of the growth of HAB species in rivers(including run-of-river impoundments), estuaries,and coastal waters, characterized by highly com-plex and stochastic mixing and flushing patterns,has remained a challenge (Thornton et al. 1990;Burkholder 2000).

    The nature of the global HAB problem in estu-arine and coastal waters has changed considerablyover the last several decades, both in extent and itspublic perception (Anderson 1989; Smayda 1990;Hallegraeff 1993). Virtually every coastal countryis now threatened by multiple harmful or toxic al-gal species, often in many locations and over broad

    areas. This trend has been referred to as the ap-parent global expansion of HABs because for manylocations, poor historic data are available. It is notclear as to how much of the increase reflectsheightened scientific awareness and scrutiny ofcoastal waters and seafood quality versus an actualincrease in the number, severity, or frequency ofoutbreaks (Anderson 1989). Many new bloom spe-cies are believed to reflect the discovery of hiddenflora populations (Smayda 1989) which had exist-ed in those waters for many years, but which hadnot been detected or recognized as harmful untilthe advent of more sensitive toxin detection meth-ods or an increase in the number and training of

    observers (e.g., Anderson et al. 1994). The num-ber of known toxic dinoflagellates has increasedfrom roughly 20 only a decade ago to at least 55today (Burkholder 1998), yet none of these morerecently known species appear to be mutants orspecies that have suddenly become toxic. Geolog-ical records or past monitoring data, where avail-able, indicate that in many locations these species

    were present in the plankton all along, but werenot discovered until recently. As underscored byHallegraeff and Bolch (1992), the accidental intro-duction of HAB species into an area via ballast wa-ter discharge can also be a contributing factor tothe global expansion.

    Of considerable concern, particularly for coastalresource managers, is the potential relationship be-tween the apparent increase in HABs and the ac-celerated eutrophication of coastal waters due tohuman activities. Linkages between HABs and eu-trophication have been noted within the past twodecades (e.g., Officer and Ryther 1980; Lam andHo 1989; Smayda 1989, 1990; Riegman 1995; Rich-ardson and Jorgensen 1996; Richardson 1997).

    Coastal waters are receiving massive and increasingquantities of industrial, agricultural, and sewage ef-fluents through a variety of pathways (Vitousek etal. 1997). In many urbanized coastal regions, theseanthropogenic inputs have altered the size andcomposition of the nutrient pool which may, in

    turn, create a more favorable nutrient environ-ment for certain HAB species.

    From innovative syntheses of available databasesworldwide, Smayda (1989, 1990) made a compel-ling case for the increase in blooms of some HABspecies being a result of coastal eutrophication. Hepresented a unifying framework that stressed anal-ogies in phytoplankton community response acrossgeographic regions and encouraged scientists andresource managers to consider the previously ne-glected role of accelerated eutrophication inHABs. Now, more than a decade later, the heavypublic and scientific attention given to HABs andapparent increasing trends, new outbreaks, or, in

    a few cases, outbreaks that have diminished in sizeor frequency, suggest that it is time to assess sci-entific progress in some of the issues that relate topossible human-induced changes in HAB distri-bution and dynamics. In particular, emphasis isneeded on the physiological, ecological, and envi-ronmental mechanisms involved. There is no ques-tion that nutrients are required by HABs, as theyare by all algal species. Here we address currentinsights into the relationships between HABs andeutrophication, focusing on sources of nutrients,the known effects of nutrient loading and reduc-tion, new understanding of pathways of nutrientacquisition among HAB species, and the specific

    relationships between nutrients and toxic algae.Through local, regional, and global examples ofthese various relationships, we offer both an as-sessment of the state of understanding, and theuncertainties that require future research efforts.

    Sources of Nutrients and their Relationshipwith HABs

    Many sources of nutrients can stimulate harmfulalgal blooms, including sewage and animal wastes,atmospheric deposition, and groundwater inflow,as well as agricultural and other fertilizer runoff.

    Yet another source is the growing aquaculture in-dustry in many coastal areas.

    Human activities have had a tremendous impacton the global cycling of nutrients in coastal sys-tems. The export of P to the oceans has increased3-fold compared to pre-industrial, pre-agriculturallevels, and N has increased even more dramatically,especially over the last 4 decades (Caraco 1995;Smil 2001). During that time, the flux of N in-creased 4-fold into the Mississippi River and morethan 10-fold into the rivers entering the North Sea

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    4/23

    HABs and Eutrophication 707

    (National Research Council 2000; Smil 2001). Hu-man activity is estimated to have increased N in-puts to the coastal waters of the northeastern Unit-ed States generally and to Chesapeake Bay specif-ically by 68-fold (Boynton et al. 1995; Howarth1998).

    Point sources generally are less important nutri-ent contributors than nonpoint sources, when con-sidered on an annual basis (National ResearchCouncil 2000). Point sources can be a majorsource of nutrients for small watersheds within, oradjacent to, major population centers. Wastewatercontributes an estimated 67% of the N inputs toLong Island Sound annually, largely due to sewagefrom New York City. Sewage treatment plants de-liver from 4080% of the N to Kaneohe Bay, Ha-

    waii, and to Narragansett Bay, Rhode Island (Nix-on and Pilson 1983; National Research Council1993). More rarely, point sources can be majorcomponents of nutrient loads to moderately sized

    watersheds. One point source, the worlds largestphosphate mine, added 50% of the total P loadingto the mostly agricultural Tar-Pamlico watershed ofthe Albemarle-Pamlico estuarine system in NorthCarolina for nearly 30 years (ca. 2,800 metric tonsof free phosphate dust added per day to the Pam-lico Estuary; reduced by 90% in the early 1990s;North Carolina Department of Environment,Health and Natural Resources [NC DEHNR]1994).

    Nonpoint sources of nutrients (from agriculturalactivities, fossil-fuel combustion, and animal feed-ing operations) are often of greater concern thanpoint sources because they are larger and more

    difficult to control. Howarth et al. (1996) estimat-ed that sewage contributes only 12% of the flux ofN from the North American continent to theNorth Atlantic Ocean. Only ca. 25% of the N andP inputs to Chesapeake Bay come from wastewatertreatment plants and other point sources (Boyntonet al. 1995). Even in relatively large watersheds theimportance of point source contributions increasesduring summer low-flow conditions, when treatedand untreated wastewater can represent 50% ormore of the river flow (e.g., the Neuse estuary ofthe Albemarle-Pamlico estuarine system; NCDEHNR 1994). This point becomes especially im-portant, given the fact that many harmful algal spe-cies are most active in summer low-flow periods.

    Fertilizer application on land remains a majorcontributor to nonpoint nutrient pollution, andthis source is still increasing at an alarming rate inmany geographic regions (Vitousek et al. 1997).Both industrial and developing nations are usingsignificantly higher loadings of fertilizer in agri-culture, with global N and P fertilizer usage in-creasing 8-fold and 3-fold, respectively, since the

    early 1960s (Constant and Sheldrick 1992; Caraco1995; Matson et al. 1997; Smil 2001). There is adirect relationship between population develop-ment, fertilizer applications, and riverine N and Pfluxes (Fig. 1a,b; Caraco 1995; Smil 2001). Whenthese nutrient supplies reach lower rivers, estuar-

    ies, and coastal waters, they are available for phy-toplankton uptake and growth. The nitrate com-ponent of fertilizers can travel long distances. Forexample, Mallin et al. (1993) demonstrated a sig-nificant relationship between nitrate, carried ca.400 km downstream to the lower Neuse estuary(over a 2-wk period), and increased phytoplanktonproductivity.

    Nutrient inputs from runoff vary not only inquantity (influenced by rainfall and other environ-mental factors), but also in composition (based onthe form of fertilizer in use), and this has impor-tant implications for HAB development. A dramat-ic trend in world fertilizer production is the in-

    creased proportion of urea in world N production,especially in third-world countries (Fig. 1c; Con-stant and Sheldrick 1992). Urea now comprisesroughly 40% of all N fertilizers produced (Con-stant and Sheldrick 1992). This is significant be-cause data indicate that in some areas, this shift infertilizer composition has resulted in a shift in thenutrient composition of runoff, potentially favor-ing some HAB species.

    Ground water has also been identified as an im-portant source of nutrients to receiving surface wa-ters. Human population growth and agriculturalpractices have increased nutrient loadings toground water, and this has the potential to affect

    algal growth in adjacent rivers, lakes, estuaries, andcoastal zones. In lakes the linkages betweengroundwater nutrient inputs and HABs (mostly ascyanobacteria) has been clearly demonstrated;

    Jones and Bachmann (1975) and Dillon and Rigler(1975) were able to reliably predict late summerphytoplankton biomass in natural lakes by takinginto account the P supplied from septic effluentleachate. In coastal areas such linkages can bemore complex and more difficult to prove conclu-sively.

    Some success has been achieved relatinggroundwater flow to the growth of the harmfulbrown tide species Aureococcus anophagefferens inLong Island, New York. This species has been as-sociated with loss of eelgrass meadows and reduc-tion in reproduction and growth of shellfish (Tra-cey 1988; Dennison et al. 1989; Gallagher et al.1989). LaRoche et al. (1997) hypothesize that inspecific coastal bays, years with high inputs ofground water lead to high dissolved inorganic ni-trogen (DIN) concentrations. A. anophagefferens isnot a strong competitor when DIN is high, as

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    5/23

    708 D. M. Anderson et al.

    Fig. 1. Nutrient inputs to the worlds oceans. A) The rela-tionship between population density in watersheds and exportof soluble reactive phosphorus (SRP) in river water, considering32 major rivers (from Caraco 1995). B) The relationship be-tween the rate of fertilizer applications and the flux of riverinenitrogen in many of the worlds coastal ecosystems (from Smil2001). C) Trends in the proportion of the contribution of ureato world N fertilizer production from 1960 to 1990 (from Con-stant and Sheldrick 1992).

    Fig. 2. Example of localized, significant increase in atmo-spheric ammonium (squares) from concentrated animal oper-ations (circles; Sampson County, North Carolina, where thereare 48 swine per person [ca. 2 million swine in total]). Approx-imately 72% of the variability in airborne ammonia during thepast decade can be explained by the expansion of the countysswine population, alone. Much of the nitrogen volatilized asammonia during spray-application of swine effluent onto fieldsis deposited into receiving rivers and streams within a distanceof ca. 100 km radius (modified from Mallin 2000).

    shown in nutrient enrichment studies in meso-cosms, where the cell density of A. anophagefferens

    was inversely correlated with DIN concentrations(Keller and Rice 1989). When groundwater inputis low, decay of the algal biomass created in pre-

    vious years from high DIN leads to elevated levelsof dissolved organic nitrogen (DON) which A. an-ophagefferens can use efficiently (Berg et al. 1997).

    A groundwater index relationship has been for-mulated that correctly hindcasts brown tideblooms in 9 of 11 years on Long Island, but therelationship has not held for all embayments in

    which this species blooms (Gobler 1999; Lomas etal. 2001; Borkman and Smayda unpublished data).

    If the groundwater hypothesis is valid, the LaRocheet al. (1997) study also suggests that there can bea significant time lag between human activities thatenrich the ground water (such as heavy fertilizerusage) and the eventual HAB impact. In Long Is-land Sound, it is possible that the massive browntides which began suddenly in 1985 may reflectheavy fertilizer usage on land 10 or 20 years earlier.

    On local to global scales, one of the most rapidlyincreasing sources of nutrients to both freshwatersand the coastal zone is the atmosphere (Figs. 2 and3). Phosphate adsorbed onto fine particulates, andnitrate derived from particulate or oxidized nitric/nitrous oxides in wet and dry deposition, have longbeen recognized as important sources of nutrientsto streams and lakes, and can be major sourcesespecially for softwater, nutrient-poor freshwatersystems (Likens et al. 1979; Kilham 1982; SwedishMinistry of Agriculture 1982). In estuarine andcoastal waters, it has been estimated that 2040%of N inputs can be of atmospheric origin, fromindustrial, agricultural, and urban sources (Duce1986; Fisher and Oppenheimer 1991; Paerl 1995,

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    6/23

    HABs and Eutrophication 709

    Fig. 3. Trends in fertilizer use and the number of red tidesreported for Chinese coastal waters (data redrawn from Smil2001 and Zhang 1994). While the general pattern is increasing

    for both parameters, it is thought that atmospheric depositionmay also play an important role in the development of theseblooms.

    1997; Driscoll et al. 2001). In other areas more re-moved from such sources, this proportion can belower, such as in the Gulf of Mexico where atmo-spheric inputs (12% of the total) are over-

    whelmed by contributions from the Mississippi andAtchafalaya Rivers (Paerl et al. 2000).

    Atmospheric inputs are important not only be-cause of their magnitude, but because the mix ofatmospheric nutrients, like other nutrient sources,can stimulate some phytoplankton species dispro-

    portionately over others. Experimental manipula-tions have shown that rainwater can enhance pro-ductivity more than the addition of a single Nsource (Paerl 1997). The high proportion of DONin rainwater, representing up to 40% of its total N,is thought to be significant in this enhancement(Timperley et al. 1985; Paerl 1997). Blooms in the

    Yellow Sea of China, which have escalated in fre-quency over the past several decades (Fig. 3), havebeen related to atmospheric deposition in additionto direct nutrient runoff (Zhang 1994). It is esti-mated that a typical rain event over the Yellow Seamay supply sufficient N, P, and Si to account for50100% of the primary production of a HABevent (Zhang 1994).

    The atmosphere, through both wet and dry de-position, may also be a source of key trace metalssuch as Fe (Church et al. 1991; Duce and Tindale1991). Phytoplankton in many estuaries and coast-al waters (where most HAB species occur) can beFe-limited (e.g., Wells 1999), and additions of at-mospheric Fe could therefore contribute to somebloom events (Martin and Fitzwater 1988; Cullen

    1991; Coale et al. 1996). Interactions between Feand N can influence plankton community struc-ture (e.g., DiTullio et al. 1993), and may be a fac-tor in the regulation of growth and encystment ofdinoflagellates (Doucette and Harrison 1991) andpossibly in the toxicity of diatoms such as Pseudo-

    nitzschia spp. (Rue and Wells unpublished data).Aquaculture ponds and cage culture systems rep-

    resent another source of nutrients, provided asfeed or fertilizer and by the biological transfor-mations occurring in these high biomass systems.It has been suggested that these enriched systemsmay promote the growth of harmful species notpreviously detected in the source water (Anderson1989; Hallegraeff 1993). The cultured animals re-tain only a fraction of their food, the rest decom-poses in the water column or settles to the bottomand decomposes, and either way, the nutrients re-leased from this decomposition can stimulate phy-toplankton growth (Cho et al. 1996; Burford 1997;

    Burford and Glibert 1999). The effect can be wors-ened if the aquaculture site is constructed in wet-lands (e.g., salt marshes or mangrove swamps) thatotherwise would serve as a sink rather than asource of nutrients to the system.

    Sakamoto (1986) calculated that nutrients re-leased from fish culture sites affect an area 39times the size of the aquaculture zone. In a qui-escent system, this sustained input could affect pro-ductivity in the area, but the extent of the nutrientimpact may diminish with higher rates of flushingby tides and currents. Recognizing the need fordilution, fish farming operations in the northwest-ern U.S. have shifted from easily accessible but

    poorly flushed bays and coves to areas with muchstronger currents resulting in a significant reduc-tion in particulate and dissolved nutrient buildupand reduced planktonic and benthic impacts (Ren-sel personal communication). Many fish farms indeveloping countries are located in shallow, easilyaccessible bays where nutrients can accumulateand stimulate algal blooms (e.g., Wu et al. 1994).Benthic nutrient regeneration of the accumulatedfeces and decomposing feed may be a significantand sustained source of nutrients in such systems.The situation in these environments was describedin harsh terms by Romdhane et al. (1998, p. 82),in referring to fish farms in Tunisian lagoons,. . . eutrophication following increased human ac-tivity in and around these lagoons influences themagnitude and frequency of toxic blooms. La-goons may function as traps for toxins or otherexudates from algae. We therefore stress that aqua-culture inside lagoons is a hazardous business.

    There is no simple generalization about the im-pacts of aquaculture operations on plankton com-munities, or specifically, on HABs, although it is

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    7/23

    710 D. M. Anderson et al.

    Fig. 4. Generalized ecosystem response to nutrient loading. At low levels of nutrient loading, the organismal response maybe rapid, but biomass changes would be few. At high rates ofnutrient loading, the physiological responses of the organism

    would be expected to be at or near saturating rates, and would

    show little increase, yet on a longer time scale, biomass wouldincrease.

    clear that in waters with a high density of aquacul-ture operations and poor flushing, the cumulativeinput of nutrients has impacts on plankton pro-ductivity. As is the case with the other sources ofnutrients to coastal waters, the increased nutrientloading will lead to increased phytoplankton pro-

    duction, but whether this leads to toxic impactsdepends on whether toxic species are present andon the relative abundance of the nutrient ele-ments, the mixing and hydrographic characteris-tics of the area, and other factors such as grazingintensity or light availability.

    Nutrient Loadings, Nutrient Reductions, andHigh-Biomass HABs

    On local, watershed, and global scales, strongcorrelations have been shown between total P in-put into freshwaters, and between total N inputinto estuaries and coastal waters and total phyto-plankton production (Schindler 1977; Wetzel

    1983; Nixon 1992; Mallin et al. 1993). HAB species,like all plant-like organisms require certain majorand minor nutrients for their nutrition, and thesecan be supplied either naturally from freshwaterand marine biogeochemical processes or throughhuman activities such as pollution. These nutrientsources include dissolved inorganic and organiccompounds of various types, as well as particulatenutrients in the form of other organisms or detri-tus.

    In attempting to understand the impacts of nu-trient availability and nutrient loading on an aquat-ic ecosystem, it is important to make the distinc-tion between effects on physiological processes or

    productivity versus biomass accumulation. As ini-tially developed conceptually by Caperon et al.(1971) and applied more recently to the Chesa-peake Bay (Malone et al. 1996), nutrient loadingresponses can be viewed in a manner analogous toa saturating response curve (Fig. 4). The effects ofnutrients may fall in the minimal response region,

    which is dominated by rapid physiological adjust-ment and low biomass accumulation, or alterna-tively, in the maximum response region, in whichphysiological processes have become saturated, butbiomass accumulations continue. The minimumresponse region of the curve also represents theperiod of bloom initiation, whereas the maximumresponse region represents bloom maintenance.

    As the period of bloom initiation is characterizedby minimal increases in biomass, the role of nutri-ents in bloom initiation is far less understood thanfor the period during which a bloom may havebeen maintained. Ultimately, the entire responsemay be saturated at exceptionally high loadingrates due to limitation by some other factor. Withinthis framework, it is important to recognize that in

    the minimum response region, impacts are few,difficult to detect, and easy to reverse, while in themaximum response region, impacts are large andoften easy to detect, but substantially more difficultto reduce and control.

    Nutrients can stimulate or enhance the impactof toxic or harmful species in several ways. At thesimplest level, harmful phytoplankton may in-crease in abundance due to nutrient enrichment,but remain in the same relative fraction of the total

    phytoplankton biomass. Even though non-HABspecies are stimulated proportionately, a modestincrease in the abundance of a HAB species cancause it to become noticeable because of its toxicor harmful effects.

    A more frequent response to nutrient enrich-ment occurs when a species or group of speciesbegins to dominate under the altered nutrient re-gime. In deeper freshwater, estuarine, and coastalmarine systems, phytoplankton dominate the algalflora. Macroalgae and benthic microalgae oftendominate many lakes and shallow, poorly flushedestuaries, lagoons, and upper embayments, as wellas coral reefs and rocky intertidal/subtidal habitats(Harlin 1993). In surface waters across the entiresalinity gradient, there are many examples of over-growth and high biomass blooms by phytoplank-ton, benthic microalgae (especially epiphytes), andmacroalgae. In many cases, the responding domi-nant species are not toxic and, in fact, are benefi-cial to coastal productivity until they exceed theassimilative capacity of the system, after which an-oxia and other adverse effects occur. When that

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    8/23

    HABs and Eutrophication 711

    threshold is reached, seemingly harmless speciescan have negative impacts.

    In this context, much has been written about thelinks between freshwater flow, nutrient loading (astotal P and phosphate), and increased nontoxic (as

    well as toxic) cyanobacterial blooms in lakes, and

    the associated bottom-water anoxia, benthic ani-mal mortalities, and fish kills that can follow theseoutbreaks (Vallentyne 1974). Freshwater flow andnutrient loading (mostly as nitrate) have beenlinked to increased numbers of estuarine algalblooms (as diatoms and other typically benign mi-croalgae or as macroalgae), followed by oxygen def-icits and finfish and/or shellfish kills (Harlin 1993;Mallin et al. 1993).

    Increases in high biomass phytoplankton bloomshave been reported from the south China Sea (Qiet al. 1993), the Black Sea (Bodeanu and Ruta1998), Hong Kong (Lam and Ho 1989), and manyother locations, typically in parallel with the nutri-

    ent enrichment of coastal waters. In ChesapeakeBay, high phytoplankton biomass is typically ob-served in the spring, associated with high riverinenutrient inputs (Glibert et al. 1995; Malone et al.1996). These large spring blooms eventually settleto the bottom, where heterotrophic bacteria pro-cess a major fraction of the organic material. Thiscan result in depletion of oxygen as temperatures

    warm (Malone et al. 1986; Shiah and Ducklow1994), leading to anoxia and benthic mortalities(e.g., Boynton et al. 1982; Malone et al. 1983; Fish-er et al. 1988, 1992; Glibert et al. 1995). As anotherexample, spring eutrophication from the N load-ing of the Mississippi and Atchafalya Rivers to the

    Gulf of Mexico has resulted in enhanced phyto-plankton production and the development of an-oxia in the Gulf of Mexico, a so-called dead zonethat has altered benthic food web dynamics sub-stantially (Turner and Rabalais 1994; Rabalais et al.1996).

    One of the clearest examples of the direct de- velopment of a toxic species in response to in-creased nutrient loading is the development ofPseudo-nitzschia spp. on the Louisiana shelf in theextended plume of the Mississippi River. BloomsofPseudo-nitzschiaspp. develop in high abundancesduring the spring when nutrient loading is highest(Dortch et al. 1997; Parsons et al. 1998, 1999; Pan2001). Both historical data and frustules preservedin cores (Dortch et al. 1997, 2000; Parsons et al.2002) indicate a large increase in Pseudo-nitzschiaspp. abundance since the 1950s, concomitant withincreases in nutrient loading. Studies in meso-cosms have also demonstrated a disproportionalincrease in Pseudo-nitzschia spp. following nutrientpulsing (Dortch et al. 2000).

    Flushing rate or turnover time (the rate at which

    all of the nutrient-laden water is exchanged ormoved out of the lake, river, or estuary) and waterdepth play a major role in the duration of the pe-riod in which nutrients are available to algal assem-blages. Lakes and reservoirs with high flushingrates and high P loading have significantly less al-

    gal production than similar systems with poorflushing (e.g., Dillon 1975; Canfield and Bach-mann 1981). The same is true of flushing in estu-aries and coastal waters, where shallow systems typ-ically support more algal growth than deeper sys-tems (Wetzel 1983; Day et al. 1989). ChesapeakeBay has an estimated mean turnover time of ca. 35d and a mean depth of ca. 9 m (Magnien et al.1992). N and P loads are estimated at ca. 80 106

    kg N yr1 and 4 106 kg P yr1, of which 5570%is delivered during the winter-spring freshet (Mag-nien et al. 1992; Boynton et al. 1995). Phytoplank-ton biomass during early spring blooms that aresupported by these nutrient supplies can exceed

    50 g chl a l1 (Glibert et al. 1995; Malone et al.1996). The Neuse estuary has a mean water turn-over time of ca. 80 d and a mean depth of ca. 3.54 m (Glasgow and Burkholder 2000; Glasgow et al.2001a). In this smaller, poorly flushed, shallow sys-tem, loadings of ca. 5 106 kg N yr1 and 68 105 kg P yr1 have supported late winter-springblooms of benign (nontoxic) dinoflagellates withbiomass as high as 300 g chl a l1 (Glasgow andBurkholder 2000; Glasgow et al. 2001a).

    Repeated incidence of increased, high-biomassblooms provide evidence of a broadly based, stim-ulatory effect on phytoplankton from anthropo-genic nutrients. The evidence for this relationship

    is further strengthened by repeated observationsthat HABs tend to decrease when nutrient loadingis reduced. Among the most cited early reports ofpartial reversal of cultural eutrophication in fresh-

    water involved removing sewage discharges fromLake Washington within metropolitan Seattle,

    Washington (Edmondson 1970). This lake had sus-tained noxious cyanobacteria blooms prior to the1920s because of raw sewage inputs. Zero dis-charge of sewage to Lake Washington was imposedin 1968, and the cyanobacterial blooms declined.In a much larger system, Great Lake Erie, thegreen macroalga Cladophora had choked much ofthe west basin with massive growth until improved

    wastewater treatment and detergent phosphatebans in the early 1980s led to significant reductionin the nuisance blooms (Ashworth 1986).

    Reduced nutrient loading similarly has promot-ed declines in estuarine and marine coastal HABs.Sewage discharges to the Mumford Cove, a shallowestuary in Connecticut were rerouted to another

    waterway in the late 1980s, and within two yearsmassive nuisance blooms of the macroalga, Ulva

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    9/23

    712 D. M. Anderson et al.

    Fig. 5. Trends in industrial production (circles) and the

    number of visible red tides (squares) in the Seto Inland Sea ofJapan. The vertical line represents the passage of the Seto In-land Sea Law in 1973, after which nutrient loadings were re-duced to one-third of 1974 levels. The triangles denote thenumber of red tides with fisheries impacts (redrawn from Okai-chi 1997 with additional data from Fukuyo).

    lactuca, were eliminated (Harlin 1993). In the SetoInland Sea in Japan between 1965 and 1976, thenumber of red tide outbreaks (high biomassblooms) increased 7-fold (Okaichi 1997), in par-allel with the increase in industrial production andchemical oxygen demand (COD) from domesticand industrial wastes (Fig. 5). In 1973, Japanese

    authorities instituted the Seto Inland Sea Law toreduce COD loadings to half of the 1974 levelsover a 3-yr period. The number of red tides beganto decrease in 1977, eventually falling to less than30% of the peak frequency, which had been in ex-cess of 300 blooms yr1 (Fig. 5). This lower levelof bloom incidence has been maintained to thepresent. These data demonstrate a general in-crease in phytoplankton abundance due to over-enrichment, and a proportional decrease inblooms when that loading was reduced. It is inter-esting that toxic blooms (in this instance, thosethat cause fish mortalities or other fisheries dam-age) also decreased after the loadings were re-duced (Fig. 5).

    Another important observation from the SetoInland Sea is that as the waters became less eutro-phic and large biomass blooms decreased, there

    was a shift in species composition, leading to agreater prevalence of some that are responsible forshellfish poisonings in humans, such as Alexan-drium tamarense and A. catenella. Paralytic shellfishpoisoning (PSP) caused by these species was not

    reported in the Inland Sea several decades ago, butis common now (Fukuyo personal communica-tion). This emphasizes a common dilemma facedby coastal managers, namely that effluent controlsmay reduce the number of phytoplankton blooms,but those actions may not result in fewer HAB im-

    pacts. This can happen because some species (andtheir high biomass blooms) may decrease in fre-quency or disappear as the waters become cleaner,but there are other harmful or toxic species thatcan fill that niche and have negative impacts. Thisreflects the great variation among HAB species inthe levels of nutrients that are optimal for growth.In some cases, oligotrophic HAB species that arenot good competitors when nutrient loads are highcan thrive as loadings from land diminish. PSP-pro-ducing Alexandrium spp. have long occurred in

    Alaska, northeastern Canada, and northern Japan;all areas with relatively unpolluted and historicallypristine waters (e.g., Horner et al. 1997). On Long

    Island, shellfishermen who have been devastatedby recurrent A. anophagefferens brown tides since1985 point out that immediately prior to the out-breaks, the affected waters are cleaner with moretransparency compared to the past when browntides did not occur (McElroy 1996). Reductions innutrients generally will reduce blooms, but maynot necessarily reduce all the potentially harmfulimpacts of HABS or all of the HAB species.

    Another example of the effect of nutrient re-duction comes from the freshwater-to-brackish Po-tomac River, a tributary of the Chesapeake Bay,

    where phosphate removal from sewage began inthe late 1970s. This region had previously experi-

    enced repeated blooms of Microcystis spp. withchlorophyll concentrations in surface waters ex-ceeding 70 g l1, but after the nutrient reduc-tions, there were sustained decreases both in totalchlorophyll and in the frequency and intensity ofthe Microcystis blooms ( Jaworski 1990). Chloro-phyll levels were generally 20 g l1.

    A final example is from the northwestern BlackSea, which experienced heavy pollution loading inthe 1970s and 1980s due to industrialization, fer-tilizer use, and urbanization in eight countries

    within that watershed, followed by reductions inthese loads in the 1990s. Significant increases ininorganic and organic nutrients were noted overthat initial 20-yr interval: nitrate was 2.58 timeshigher, and phosphate was up to 20-fold higher(Bodeanu 1993). A consequence of this enrich-ment was an increase in the frequency and mag-nitude of algal blooms, as well as changes in thespecies composition. In the 1960s, high biomassblooms were rare, but during the two decades ofintense eutrophication pressure, blooms becamerecurrent, with cell densities greatly exceeding past

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    10/23

    HABs and Eutrophication 713

    abundance levels (Bodeanu 1993). During the1980s, when nutrient loadings peaked, 49 majorblooms were reported, of which 15 had 10 mil-lion cells l1 (Bodeanu and Ruta 1998). Anoxia,fish mortalities, and other impacts were frequent.

    A characteristic of this interval was the decreased

    abundance of diatoms and larger algae, and theirreplacement by flagellates and nanoplankton. In astriking reversal, algal blooms began to decrease in1991, both in number and size, and this trend hascontinued to the present. Diatoms became moredominant, and nanoplankton and flagellates de-creased. From 19911996, there were only threeblooms with cell concentrations in excess of 10 mil-lion cells l1. This reduction in blooms coincided

    with significantly decreased fertilizer usage as a re-sult of the loss of economic subsidies that accom-panied the breakup of the former Soviet Union(Bodeanu and Ruta 1998). It will be interesting tosee if the positive trend in bloom incidence of re-

    cent years is reversed when economic develop-ment, and thus fertilizer usage, increase in thecoming years.

    There are a number of examples where increas-es and decreases in nutrient loadings due to hu-man activities have resulted in parallel increases ordecreases in bloom incidence. Many of these ex-amples are of high biomass blooms, that causeharm through excessive population developmentand its decay. Other factors need to be consideredin understanding phytoplankton compositionalchanges that lead to development of HAB out-breaks, but not necessarily to high biomass pro-duction.

    Nutrient Composition and HAB Development

    Many factors affect phytoplankton species com-position and bloom development, and amongthese is the composition of the nutrient pooltheforms of the nutrients supplied, as well as the rel-ative abundance of the major nutrient elements.Some generalities are beginning to emerge withrespect to the preference of many bloom-formingspecies for specific forms of nutrients, as well asthe tendency for some blooms to occur when theratios of nutrient availability or supply are altered.The latter concept is based largely on the nutrientratio hypothesis (Tilman 1977; Smayda 1990, 1997)

    which argues that environmental selection of phy-toplankton species is associated with the relativeavailability of specific nutrients in coastal waters,and that human activities have altered these nutri-ent supply ratios in ways that change the naturalphytoplankton community composition and possi-bly favor harmful or potentially toxic forms.

    Perhaps the clearest demonstration of the effectof altered nutrient supply ratios involves the stim-

    ulation of non-diatom species following changes inthe availability of N or P relative to silicate. Dia-toms, the vast majority of which are harmless, re-quire silica in their cell walls, whereas most otherphytoplankton do not. Since silica is not abundantin sewage effluent but N and P are, the N:Si or P:

    Si ratios in some lakes, rivers, estuaries, and coastalwaters have increased over the last several decades(Shelske et al. 1986; Smayda 1989, 1990; Rabalaiset al. 1996). In theory, diatom growth will cease

    when silica supplies are depleted, but other phy-toplankton classes can continue to proliferate us-ing the excess N and P.

    Research is ongoing in various geographic re-gions to further examine this concept, which issupported by several data sets. From a long-termdatabase in Great Lake Michigan, Schelske et al.(1986) found evidence of silica depletion that wascorrelated with increased anthropogenic P loadingthrough the early 1970s. By the 1980s, cyanobac-

    teria and colonial green algae had increased to co-dominance with diatoms, but at that point P inputsbegan to decline. The phytoplankton communitythen shifted from ca. 50% cyanobacteria and co-lonial greens to replacement by flagellates in sum-mer with diatoms dominant in the spring. Similar-ly, in marine waters of Tolo Harbor in Hong Kong,there was an 8-fold increase in the number of redtides (mainly dinoflagellates) per year between1976 and 1989, in parallel with a 6-fold increase inhuman population density and a 2.5-fold increasein nutrient loading in that watershed that alteredthe nutrient ratios (Lam and Ho 1989). In the midto late 1980s, as pollution loadings decreased due

    to the diversion of sewage effluent to Victoria Har-bor, there was a resurgence of diatoms and a de-crease in dinoflagellates and red tides (Yung et al.1997).

    These blooms in Tolo Harbor show a distinct re-lationship with nutrient ratios, but not just N:Si orP:Si. Hodgkiss and Ho (1997) demonstrated thatthe numbers of dinoflagellate red tides increasedas the annually averaged N:P ratio fell from 20:1to 11:1 between 1982 and 1989 (Fig. 6). In moredetailed analysis of the patterns during a single

    year, Hodgkiss (2001) showed that whenever theN:P ratio fell below 10:1 in Tolo Harbor, dinofla-gellate cell numbers increased. These two inversecorrelations are consistent with experimental data,

    whereby the three major dinoflagellate species inTolo Harbor in the 1980s (Prorocentrum micans, P.sigmoides, and P. triestinum) were shown to have op-timal N:P ratios for growth of 510, 415, and 815:1, respectively, all significantly below Redfieldproportions. As the N:P ratio in Tolo Harbor de-creased between 1982 and 1989, these species in-creased in abundance.

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    11/23

    714 D. M. Anderson et al.

    Fig. 6. Trends in the N:P molar ratio (circles) and the num-ber of reported red tides (squares) in Tolo Harbor, Hong Kong

    from 1980 to 1990 (redrawn from Hodgkiss and Ho 1997).

    Fig. 7. Change in N:P molar ratio (circles) in Dutch coastal waters coincident with increase in Phaeocystis blooms (squares;redrawn from Riegman 1995).

    In Tunisian lagoons where Gymnodinium aureo-lumi (formerly Gyrodinium aureolum) was found tobe the cause of repeated fish kills in aquaculturesystems, blooms occurred when the N:P ratio(which was normally very high) began to declinein the autumn (Romdhane et al. 1998). There isevidence to suggest that the ichthyotoxic dinofla-gellate Pfiesteria piscicidamay do disproportionately

    well when the ratio of N:P decreases following anincrease in the availability of phosphate (Burk-holder and Glasgow 1997; Burkholder et al. 2001b;

    Glasgow et al. 2001b). Another prominent example of the importance

    of nutrient supply ratios in determining phyto-plankton species composition is seen with thefoam-producing prymnesiophyte Phaeocystis pouch-eti. A 23-year time series off the German coast doc-uments the general enrichment of these coastal wa-ters with N and phosphate and a 4-fold increase inthe N:Si and P:Si ratios (Radach et al. 1990). This

    was accompanied by a decrease in the diatom com-munity and an increase in the occurrence ofPhaeo-cystis blooms. Mass occurrences of this species be-gan in 1977 in the North Sea (Cadee and Hege-man 1986) and increased in cell abundance andbloom duration through 1985. The general N andP enrichment of that coastal area resulted in winterconcentrations an order of magnitude higher thanthose in adjacent Atlantic waters (Lancelot 1995).The abundance of these nutrients is less of an issuethan their relative proportions. These blooms werefirst related to the increase in N:Si ratios, particu-larly following the spring diatom blooms which de-pleted the silica but not the nitrate (Cadee and

    Hegeman 1986; Smayda 1990). Riegman (1995)further showed that in mixed phytoplankton as-semblages in the laboratory, P. poucheti becamedominant only when N:P ratios were 7.5 or lower,and at N:P ratios of 1.5, there was almost completeP. poucheti dominance. These relationships are con-sistent with the trends for summer blooms of P.

    poucheti in Dutch coastal waters, which accompa-nied a shift from P-limitation to N-limitation in thearea; lower N:P ratios coincided with higher, andmore variable, P. poucheti abundance (Fig. 7).

    Nutrient ratios may also be affected by other

    types of human development in addition to directnutrient pollution. The building of dams has nu-merous associated environmental problems, in-cluding the potential for altered water quality. Damconstruction, coincident with increased P loading,has led to diatom blooms, and thus to the seques-tration of silica (Turner and Rabalais 1991). In thedevelopment of the massive Three Gorges Dam inthe upstream region of the Changjiang (YangtzeRiver), the potential for eutrophication and othermassive environmental and cultural damage hasbeen greatly debated (Zhang et al. 1999). In thissystem, it is thought that by the year 2010, silica

    will be significantly reduced due to diatom uptakeand sediment trapping by the dam, and this com-bined with the trend of increasing N loading willlead to very high N:Si and N:P ratios downstream.

    As the Changjiang watershed supplies nearly 10%of the total world populations water resources and40% of the Chinese national food production, thesocietal benefits from the dam are significant, as isthe potential for negative impacts on the health ofcoastal ecosystems (Zhang et al. 1999).

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    12/23

    HABs and Eutrophication 715

    Fig. 8. Relationship between DOC:DON for numerousbloom periods and locations. In each case, the dark bars rep-resent either long-term average non-bloom periods for the samesites or comparable sites outside of the bloom region. The graybars represent the periods during the peak of bloom occur-rence. The non-bloom data from the P. piscicida and P. minimumblooms represent long-term monitoring results for tributaries ofChesapeake Bay. The P. piscicida bloom occurred in these sametributaries in Maryland in 1997, and the P. minimum bloom oc-curred in the same region in 1998. The A. anophagefferensblooms were sampled either in New York or Maryland coastalbays in 1999. The Gymnodiniumspp. bloom was sampled duringa red tide event in Kuwait Bay, 1999 (data were derived andredrawn from Glibert et al. 2001; Lomas et al. 2001; and Glibertunpublished data).

    The nutrient ratio concept has recently been ex-panded to include the relative abundance of dif-ferent chemical forms of nutrients, such as organic

    versus inorganic N and carbon (C) compounds.Recent studies in enriched coastal areas haveshown that while productivity may increase quan-

    titatively with overall N availability, the DON com-ponent may contribute disproportionately to thechanges in phytoplankton succession, apparentlyfavoring the development of some HABs (Paerl1988; Berg et al. 1997; LaRoche et al. 1997; Lomaset al. 2001). The DON pool is composed of a widerange of compounds from small amino acids andurea to complex molecules such as proteins andhumic acids. Some are available for assimilation bythe phytoplankton, whereas many other com-pounds are highly refractory and not readily used.One component of the DON pool, urea, has beenshown to be highly correlated with the outbreak ofharmful dinoflagellates in estuarine fish ponds

    (Glibert and Terlizzi 1999), where elevated levelsof urea were associated with significant dinoflagel-late outbreaks 73% of the time, but urea concen-trations of 1.5 M were not associated with anydinoflagellate blooms. In several Chesapeake Baytributaries, high urea concentrations have alsobeen found to precede large blooms of the dino-flagellate Prorocentrum minimum (Glibert et al.2001). The trend toward increasing applications ofurea fertilizer (Constant and Sheldrick 1992) mayincrease the likelihood of blooms of organismsthat grow well on this nutrient.

    Several HABs have been shown to be related toan elevation in the ratio of dissolved organic car-

    bon (DOC):DON. Three separate blooms in Ches-apeake Bay occurring over a 3-yr period, includingP. piscicida, P. minimum, and A. anophagefferens, wereall correlated with elevated DOC:DON ratios rela-tive to the long-term mean (Glibert et al. 2001; Fig.8). The elevation in this ratio for these particularblooms was a reflection of both elevated levels ofDOC as well as a depletion of DON. Lomas et al.(2001) have shown this relationship to be robustfor numerous brown tide blooms in Long Island,New York (Fig. 8). During a bloom of Gymnodiniumspp. in Kuwait Bay, the ratios of DOC:DON for sta-tions collected within the bloom were approxi-mately twice those determined for non-bloom sta-tions with a mixed phytoplankton assemblage(Heil et al. 2001; Fig. 8). This relationship is de-serving of additional study in other bloom condi-tions. Of particular interest in this context is thepotential change in DOC:DON preceding blooms.

    Pathways of Nutrient Acquisition

    An understanding of physiological responses isfurther complicated by the fact that the rate of nu-

    trient supply will not necessarily correlate with therate of nutrient assimilation by the algae, as thelatter is controlled by nutritional preferences, up-take capabilities, and physiological or nutritionalstatus. The response by either the total phytoplank-ton community or individual species within thecommunity also depends on many factors, includ-ing interactions with grazers and physical forcingssuch as turbulence. Grazers may inhibit the devel-

    opment of phytoplankton biomass through theirfeeding, while at the same time, enhance the re-generation of nutrients through their release andexcretion. This in turn will alter the balance of re-duced versus oxidized forms of N (Glibert 1998).

    The assimilation of nutrients by phytoplanktondepends on environmental factors such as light,temperature, and water column stability with dif-ferent environmental effects having differential im-pacts on different nutrient substrates. The uptakeof ammonium and urea are usually thought to beless light dependent than the uptake of nitrate(MacIsaac and Dugdale 1972; Fisher et al. 1982),and the temperature dependence of ammoniumuptake may also differ from that of nitrate (Lomasand Glibert 1999a). Water column stability is an-other critical factor influencing species composi-tion. Blooms of Karenia cf. mikimotoi have been as-sociated with warm, stable conditions, and can per-sist for extended periods with low light and lownutrients (Dahl and Tangen 1993). In Norwegian

    waters, these blooms initiate at the pycnocline inthe summer or early autumn in offshore waters,

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    13/23

    716 D. M. Anderson et al.

    then collect at hydrographic fronts in nearshore waters (Dahl and Tangen 1993). In Tunisian la-goons, where blooms of G. aureolum have causedrepeated fish kills, a correlation has been foundbetween the development of blooms and decreas-ing day length, consistent with the frequency of

    these blooms being greater in late summer or au-tumn (Romdhane et al. 1998). Any potential ef-fects of nutrient stimulation on HAB biomass orproductivity must be considered within the physi-cal and environmental tolerances of the particularspecies of concern.

    In recent years, the physiological strategies bywhich different groups of species acquire their nu-trients have become better understood. Rapidlygrowing marine diatoms have been highly corre-lated with large and/or frequent additions of ni-trate, in part because they have physiological ad-aptations which allow them to exploit nitrate-richconditions (Takahashi et al. 1982; Goldman 1993;

    Lomas and Glibert 1999a,b, 2000). Microflagella-tes, including dinoflagellates, are most frequentlyassociated with low nitrate concentrations, higherammonium, urea, or DON supply, and consistentphysiological preference for reduced N forms(e.g., Berg et al. 1997; Carlsson et al. 1998; Lomasand Glibert 1999b). Most estuarine and coastal ma-rine HAB species are microflagellates. Harmful es-tuarine dinoflagellates tend to occur in waters thathave seasonally high phosphate and nitrate, as wellas high DOC and other organic nutrient forms(Burkholder and Glasgow 1997; Burkholder et al.1997, 2001a,b; Magnien et al. 2000; Glasgow et al.2001a; Glibert et al. 2001). Indeed, the brown tide

    species that blooms in Texas, Aureoumbra lagunensisis incapable of nitrate uptake, and thus must usereduced N forms (DeYoe and Suttle 1994).

    An important physiological adaptation of manyflagellate species, including some HAB species, isthe ability to acquire both N and C via particleingestion or by the uptake of dissolved organiccompounds (reviewed in Graneli and Carlsson1998). Such mixotrophic or heterotrophic tenden-cies have been linked with the ability of these cellsto thrive in environments where inorganic nutri-ents or light may otherwise be insufficient to meettheir nutritional or C demands. Toxic Chrysochro-mulina polylepis cultures have been shown to con-sume more algal food when limited by P comparedto nutrient-replete and N-limited conditions(LeGrand et al. 1996). Mixotrophy is now consid-ered essential for the survival and growth of many

    Dinophysis species, including those responsible fordiarrhetic shellfish poisoning (DSP). This is sup-ported by uptake of 14C in the dark, either fromdirect ingestion of labelled algal prey or dissolvedorganic substances released by those algae (Gra-

    neli et al. 1997). Using different methods, Jacob-son and Anderson (1996) found food vacuolescontaining prey fragments (probably ciliates) in Di-nophysis norvegica and D. acuminata, confirmingthese species ability to ingest particulate food.Other common HAB species have also been shown

    to be mixotrophic, including Heterosigma carterae( H. akashiwo), A. tamarense (Nygaard and Tobie-sen 1993), and Gyrodinium galatheanum ( Karlo-dinium micrum; Li et al. 2000, 2001). Given the im-portance of mixotrophy in many species, as well asthe development of new methods to measure in-gestion and C uptake (Schnepf and Elbrachter1992; Stoecker 1999; Stickney et al. 2000), thenumber of HAB species known to be mixotrophic

    will likely increase as more are examined for thischaracteristic (Burkholder and Glasgow 1995,1997; Burkholder et al. 2001b).

    A unique example of mixotrophic nutrition isthe toxic Pfiesteria complex (two speciesP. pisci-

    cida and Pfiesteria shumwayae; Burkholder et al.2001a,b). These dinoflagellates are heterotrophs,

    yet they can be stimulated directly and indirectlyby inorganic as well as organic nutrient enrich-ment (Burkholder and Glasgow 1997, 2001; Burk-holder et al. 1998a, 2001a,b; Glasgow et al. 2001b;Parrow et al. 2001). Like other heterotrophic di-noflagellates (Schnepf and Elbrachter 1992), theycan take up inorganic and organic nutrients di-rectly (e.g., dissolved amino acids: Burkholder andGlasgow 1997; Glasgow et al. 1998; nitrate, am-monium, and urea: Lewitus et al. 1999a). Pfiesteriaspp. are not capable of photosynthesis on theirown, but zoospores can retain chloroplasts from

    algal prey (Burkholder and Glasgow 1997; Lewituset al. 1999a,b; Glasgow et al. 2001c). This phenom-enon, kleptochloroplastidy, is increasingly recog-nized in dinoflagellates and some protozoan cili-ates (Stoecker 1998; Skovgaard 1998).

    Kleptochloroplastidy allows Pfiesteria spp. tofunction as mixotrophs for hours to days (Lewituset al. 1999a). In this mode, cells can take up Ndirectly (Lewitus et al. 1999a). Pfiesteria spp. havealso been shown to be stimulated indirectly by nu-trient enrichment, mediated through the abun-dance of algal prey that they consume when fishare not present (Burkholder and Glasgow 1995,1997; Glasgow et al. 1998; Parrow et al. 2001). Theability to consume an array of prey ranging frombacteria to mammalian tissues, as well as dissolvedsubstrates, allows Pfiesteriaspp. to thrive where foodis abundant (Burkholder and Glasgow 1995, 2001;Burkholder et al. 2001b). Toxic Pfiesteriaoutbreakshave occurred in shallow, poorly flushed estuariesthat have been highly impacted by nutrient over-enrichment, including the Neuse, Pamlico, andNew River estuaries of North Carolina and the trib-

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    14/23

    HABs and Eutrophication 717

    utaries of Marylands eastern shore (Burkholder etal. 1995, 1997; Lewitus et al. 1995; Burkholder andGlasgow 1997; Glasgow et al. 2001a). In both lab-oratory and field studies, Pfiesteria zoospore pro-duction has been shown to be stimulated by hu-man and animal wastes (Burkholder and Glasgow

    1997, 2001; Burkholder et al. 1997). Nutrients pro-vide a food-rich habitat for Pfiesteriaspp., but otherenvironmental conditions are required for toxicPfiesteria activity, especially poor flushing, fish inabundance, and brackish salinities (Burkholderand Glasgow 1997; Glasgow et al. 2001a). The abil-ity of these heterotrophic dinoflagellates to func-tion photosynthetically, and to switch betweenmodes of nutrition and among an array of preytypes as conditions change, represents a significantsurvival mechanism.

    Many phytoplankton have the ability to acquiresome of their nutrients via extracellular oxidationor hydrolysis. Extracellular amino acid oxidation

    has been shown to occur in a wide range of fla-gellates and in a range of ecosystems, although thisprocess appears to be expressed to a greater de-gree when ambient inorganic nutrient levels are ator near depletion (Palenik and Morel 1990a,b;Pantoja and Lee 1994; Mulholland et al. 1998).Proteins and peptides may also be hydrolyzed atthe cell surface, producing smaller compoundsthat can be taken up by the cells (Hollibaugh and

    Azam 1983; Keil and Kirchman 1992; Pantoja et al.1997; Pantoja and Lee 1999). While much is stillto be learned about the role of this process in thedevelopment of HABs, there is some evidence thatcertain HAB species possess this ability (Mulhol-

    land et al. 2000).The uptake of organic compounds may contrib-

    ute to the C requirements of HAB cells, in additionto their N or P requirements. The suggestion thatC acquisition may stimulate algal growth ratesthrough organic uptake is by no means new(Schell 1974; Wheeler et al. 1974; Lewitus andKana 1994). Specific examples of the linkage be-tween DOC uptake and HAB development, how-ever, are only now beginning to emerge. In 1998,a new species of dinoflagellate, Kryptoperidiniumcarolinium (sp. ined.; formal description ongoingby Lewitus unpublished data), was observed in the

    waters of coastal South Carolina. Following inten-sive monitoring of all forms of inorganic and or-ganic nutrients, it was concluded that bloom initi-ation followed the pulsed delivery of organic nu-trients (Lewitus et al. 2001). Bloom development

    was coincident with a greater than 3-fold decreasein both DOC and DON. These findings under-score the need to incorporate organic nutrientsand heterotrophic potential in both monitoringand models of HAB population dynamics.

    Indirect Nutrient Linkages with HABs

    All too frequently, public perception of whethernutrient over-enrichment has reached undesirablelevels has been based on acute, obvious or easilymeasured symptoms, such as high biomass algalblooms, massive fish kills, and oxygen deficits. Be-cause of this focus, a broad array of indirect,chronic, often-subtle but serious impacts of nutri-ent pollution on aquatic ecosystems remain under-emphasized and, in some cases, poorly under-stood. The available data indicate that these chron-ic, indirect impacts can be important in controllingthe growth of HAB species over the long term inlakes, rivers, estuaries, and marine coastal waters.

    As eutrophication progresses, for example, shiftsin phytoplankton communities toward declines incertain diatom species in favor of less desirable na-noplankton and flagellates can lead to subtle butimportant changes at higher trophic levels. Some

    freshwater diatom species that grow best in low nu-trient regimes produce lipids that are essential forzooplankton sexual reproduction. Under nutrientover-enrichment, these species are replaced by spe-cies that produce low or negligible quantities ofthese lipids (Kilham et al. 1997). In estuarine wa-ters, spawning of green sea urchins and blue mus-sels appears to be triggered by a heat-stable metab-olite that is released in high abundance by certainspecies of phytoplankton that decline with culturaleutrophication (Starr et al. 1990). Replacementspecies that thrive under nutrient enrichment pro-duce low or negligible quantities of the substance.

    At the same time, excessive nutrient loading has

    led to the decline and, eventually, the disappear-ance of rooted vegetation that is critically impor-tant to the survival of animals such as certain zoo-plankton, finfish, and/or shellfish which graze onalgae. Overfishing has led to significant declines insome shellfish species, such as oyster populationsin Chesapeake Bay (Newell 1988; Rothschild et al.1994). Such factors would interact in depressinggrazing activity which, in turn, would indirectly en-courage growth of phytoplankton, including HABspecies, under nutrient enrichment (Burkholder2000).

    Nutrient loading seldom occurs alone. Atmo-spheric deposition contains nutrients as well as

    acid-imparting contaminants and toxic substancessuch as pesticides; cropland runoff carries not onlynutrients, but pesticides and suspended sediments(Miller 2000). Nutrients in poorly treated humansewage and animal wastes are added to surface wa-ters along with heavy metals and other toxic sub-stances, suspended solids, estrogens and estrogen-mimic substances, and a wide array of microbialpathogens (Burkholder et al. 1997; Mallin 2000;

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    15/23

    718 D. M. Anderson et al.

    Miller 2000). Excessive nutrients act in concert with these other, co-associated pollutants to causephysiological stress and disease in sensitive grazingfauna which, again, could indirectly help to pro-mote the growth of harmful algae through loweredgrazing pressure and facilitated access to weakened

    fish by some harmful algae that consume them asprey.

    Other factors such as suspended sediments orgrazing pressure may reduce or negate a poten-tially stimulatory nutrient effect. In turbid lakesand reservoirs with high episodic sediment load-ing, and systems with relatively rapid flushing rates,high P loading may not stimulate phytoplanktonblooms because of light limitation and short waterturnover times (Dillon 1975; Cuker et al. 1990;Burkholder et al. 1998b). Cyanobacteria canbloom under low light availability by taking advan-tage of periods between episodic sediment loadingevents when the water clears, or by using mecha-

    nisms for buoyancy regulation to position them-selves near the water surface (Burkholder et al.1998b; Dortch unpublished data). In lakes with lowto moderate nutrient loading, grazing pressurefrom large-bodied zooplankton can significantly re-duce the populations of most phytoplankton spe-cies, balancing the nutrient stimulation effect(Harper 1992).

    Similar observations have been reported in es-tuaries and coastal waters. The Pearl River estuarysupplies a huge pollution load to the waters of thesouth China Sea, including the western waters ofHong Kong, yet the number of red tides and gen-eral chlorophyll levels are low compared to the

    conditions in Victoria Harbor and areas to the east.Tang et al. (2001) hypothesized that this inverserelationship between nutrient loading and algalbiomass is due to the high sediment loads that ac-company the Pearl River discharge. Light limita-tion would prevent the full utilization of the nutri-ents supplied to the phytoplankton.

    In San Francisco Bay, increased nutrient loadshave resulted in increased secondary productionin the benthos, which in turn modulates the algalbiomass (Cloern 1982). In an analogous manner,primary production in the Bay of Brest, France, isnutrient limited, even with large nutrient loadingsfrom its tributaries. Nutrient inputs have increased3-fold since 1975, yet chlorophyll levels have notchanged significantly (Le Pape et al. 1996; Le Papeand Menesguen 1997). Primary productivity has in-creased slightly, but grazing pressure has as well,particularly in the benthos. In this case, the maincontrol of eutrophication pressures appears to re-late to a strong tidal influence and hydrodynamicexchange. The resulting stirring hinders the for-mation of a persistent surface mixed layer where

    phytoplankton have access to nutrient inputs andto light. Horizontal tidal currents cause significant

    water exchange with the Iroise Sea, and reduce theaccumulation of nutrients and plankton in the Bay.

    As has been observed in certain other systems, nu-trient loading has been beneficial in that it sup-

    ports increased productivity. Such beneficial effectsshould continue as long as those loadings fall with-in the assimilative capacity of the system.

    In some cases, indirect relationships between nu-trient loading or availability and the developmentof a HAB species may be difficult to establish, dueto the complexities of the nutrient cycling path-

    ways involved. These may be on short temporaland spatial scales, or on longer-term scales. Oneexample of such pathways potentially leading toHAB development involves the release of DON fol-lowing N fixation. Blooms of the N-fixing cyano-bacterium Trichodesmium have been found to re-lease a significant fraction of their newly fixed N

    in the form of ammonium and DON (Capone etal. 1994; Glibert and Bronk 1994). In denseblooms of this organism, the concentration of re-duced N forms can be enriched several-fold overcontrol sites (Karl et al. 1992; Glibert and ONeil1999; ONeil et al. submitted). It has been sug-gested that this production of reduced N fuels redtide blooms ofKarenia brevis( Gymnodinium breve)off the coast of Florida (Walsh and Steidinger2001; Lenes et al. 2001). Likewise, DON release byTrichodesmium has been shown to be correlated

    with an increase in the development of dinoflagel-lates such as Dinophysis off the coast of Australia(ONeil et al. submitted).

    Another example of indirect stimulation of HABspecies by nutrients is the ichthyotoxic dinoflagel-late, P. piscicida. In toxic strains of this organism,temporarily nontoxic zoospores are the precursorsof actively toxic zoospores. These nontoxic zoo-spores have been found to increase in response toelevations in chlorophyll (Burkholder and Glasgow1997; Glasgow et al. 2001a), and their growth rateshave been shown to vary widely depending on theform of algal prey (Burkholder and Glasgow 1995;Glasgow et al. 1998; Burkholder et al. 2001a; Par-row et al. 2001). Nutrients may select for certainphytoplankton species which may promote Pfiester-ia growth in temporarily nontoxic mode.

    Links between Nutrients and Toxicity

    The discussion thus far has centered on nutrientpools as they affect the growth and accumulationof HAB cells. There is evidence that nutrients canplay a major role in the regulation of toxicity insome HAB species, and this can have significantimplications to toxin monitoring programs andpublic health decisions. In some cases, toxicity can

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    16/23

    HABs and Eutrophication 719

    increase or decrease dramatically depending onthe limiting nutrient. Saxitoxin production by A.tamarense can be 510-fold higher in P-limited ver-sus N-limited cells (Boyer et al. 1987; Anderson etal. 1990). Likewise, domoic acid production byPseudo-nitzschia multiseries is inversely correlated

    with the ambient Si concentration in batch culture(Pan et al. 1996a). In that study, cells began accu-mulating this toxin only when the division rate de-clined as a result of partial or total depletion ofsilica. When cultures were N-limited, no toxin wasproduced. Toxin production was greatly enhancedunder P-deficient conditions in continuous cul-tures (Pan et al. 1996b). Recent results also suggestthat Fe limitation can enhance toxicity in Pseudo-nitzschia spp. (Rue and Wells unpublished data).

    For other HAB species a similar picture emerges:toxin production varies significantly with differentdegrees and types of nutrient limitation. The di-noflagellate D. acuminata produced elevated levels

    of the DSP toxin, okadaic acid, under both N andP limitation, but the enhancement was 6-fold larg-er with N-limitation ( Johansson et al. 1996). In ananalogous although opposite manner, Chrysochro-mulina polylepis was 6-fold more toxic under P en-richment than N-limited conditions ( Johanssonand Gran eli 1999a). An other pr ymnesiophy te,Prymnesium parvum, increased toxicity under N-lim-ited or P-limited conditions (Johansson and Gra-neli 1999b).

    The chemical form of the nutrient supplied tothe HAB species can also affect toxicity, althoughthis is an area that has received relatively littlestudy. K. brevis has been shown to increase its pro-

    duction of brevetoxin up to 6-fold when exposedto elevated urea levels of 0.5 to 1.0 mM in batchculture compared to controls without urea enrich-ment (Shimizu et al. 1993). The urea levels usedin that experiment far exceed those found undernatural conditions, but the implication is that cer-tain compounds are more readily assimilated andincorporated into algal toxins than others. Withthe addition of urea or glycine, the cells switchedfrom autrotrophic to heterotrophic nutrition, us-ing the C skeleton only after the N was used. Inthis study, toxicity was not influenced by the addi-tion of leucine or aspartic acid (Shimizu et al.1993).

    The ecological implications of nutrient effectson toxicity are significant. What is not yet clear ishow often the conditions that induce these chang-es actually occur in natural waters, and how humanactivities, and specifically eutrophication, affectoverall toxin potential. One can envision severalscenarios for eutrophic waters, depending on theextent of nutrient enrichment, the resulting nutri-ent availability ratios, and the HAB species and tox-

    in involved. Due to the nutrient enrichment, HABcells might be more abundant, but because of thealtered nutrient ratios, their cellular toxicity couldbe higher or lower than with non-eutrophic con-ditions. Depending on the species, the net effectcould thus be an increase, decrease, or no change

    in overall toxicity from a public health, fisheries,or ecosystem impact perspective. This is an area ofobvious importance, but further research is need-ed before useful insights about nutrient form andHABs can be provided to coastal resource manag-ers.

    HABs with Little Apparent Link toNutrient Enrichment

    A common assumption by the public and thepress is that new or unusual HAB events are some-how linked to pollution, and that all nutrient in-creases will result in algal blooms. The situation isfar from that simple, but in many cases a link be-

    tween blooms and eutrophication can be identi-fied. It should be emphasized though that thereare HABs that do not appear to have this linkage.These are blooms for which there may be no nu-trient relationship, or one that has not yet beenidentified. There may be other factors that exertmore control in regulating plankton communitydynamics. This is true for some new outbreaks andfor expansions of recognized or recurrent blooms.PSP toxicity from toxic Alexandriumspecies is a pre-sent-day problem in the relatively pristine waters ofthe Gulf of Maine, as well as along most of the U.S.

    west coast including Alaska. The blooms that occurundoubtedly use some nutrients that derive from

    human activities, given their proximity to the coast,but other factors seem to better explain the recentspreading of these organisms. The PSP problemhas expanded into southern New England and intoPuget Sound on the U.S. west coast over the lastseveral decades, but these increases are thought toreflect the transport of cyst-forming Alexandriumspecies into those regions by natural storms andcurrents and with the deposition of cysts that haveallowed the species to colonize the areas (e.g., Ren-sel 1993; Anderson et al. 1994). For Alexandriumspp. in the Gulf of Maine, increased nutrient load-ing and composition appear to be secondary fac-tors influencing growth.

    Conclusions and Cautions

    Eutrophication is a global problem, and coastalareas throughout the world have been affected.There is little question that nutrient loading fuelshigh biomass algal blooms, and increases in chlo-rophyll have been shown to parallel increases innutrient concentrations. There is clear evidencefor direct stimulation of some HABs by nutrient

  • 8/3/2019 Harmful Algal Blooms and Eutrophication Nutrient Sources, Composition, And Consequences

    17/23

    720 D. M. Anderson et al.

    over-enrichment. The linkages between otherHABs and eutrophication, however, are more com-plex and include indirect as well as direct path-

    ways; and linkages between some oligotrophicHAB species and eutrophication are not known.There have been many significant advances in our

    understanding of the physiological requirementsfor, and the mechanisms of nutrient acquisition by,HAB species. We have gained much knowledge ofhow certain nutrients, and their proportions, canregulate some species or groups of species.

    It is important to recognize that the impacts ofnutrient loading depend on many factors, from thespecies composition and nutritional state of the or-ganisms at the time of the loading, to the physicalfeatures of the environment at that point in time,as well as the existence of grazers. Similar nutrientloads will not necessarily have the same effect ona different environment, or on the same environ-ment at a different point in time. It is importantto avoid ascribing the apparent global increase inHABs solely to pollution or eutrophication, al-though the public and the press often assume thislinkage. There are many causes for the expansionand eutrophication is but one of these mecha-nisms.

    Although there have been many successes in re-lating nutrient quantity and composition to out-breaks of HABs, in general the relationships be-tween nutrient delivery and the development ofblooms of many HAB species, and between nutri-ent enrichment and the potential toxicity ofblooms or outbreaks of those species, remain poor-

    ly understood. Local, regional, and worldwide co-ordinated efforts, particularly those targeting com-parative ecosystems that include both highly eutro-phic waters and those that have experienced al-tered nutrient inputs will be required to betterunderstand the underlying direct and indirectmechanisms that interact to control the complex-ities of these relationships.

    ACKNOWLEDGMENTS

    This work was supported in part by National Oceanic andAtmospheric Administration (NOAA) Grants No. NA96OP0099,NA860P0493, and NA860P0510; NOAA Sea Grant NA86RG0075(Project R/B-158); National Science Foundation (NSF) Grants

    No. OCE-9808173, OCE-9415536, and OCE-9912089; and U.S.Environmental Protection Agency (EPA) grant No. R-825551-01-0l. This effort was supported by the U.S. Ecology and Ocean-ography of Harmful Algal Blooms (ECOHAB) Program spon-sored by NOAA, the U.S. EPA, NSF, National Aeronautics andSpace Administration (NASA), and Office of Naval Research(ONR). This is contribution number 10398 from the WoodsHole Oceanographic Institution, 3516 from the University ofMaryland Center for Environmental Science, contributionCAAE-095 from the North Carolina State University Center for

    Applied Aquatic Ecology, and 36 from the ECOHAB program.

    LITERATURE CITED

    ANDERSON, D. M. 1989. Toxic algal blooms and red tides: A glob-al perspective, p. 1116. In T. Okaichi, D. M. Anderson, andT. Nemoto (eds.), Red Tides: Biology, Environmental Scienceand Toxicology. Elsevier, New York.

    ANDERSON, D. M., D. M. KULIS, G. J. DOUCETTE, J. C. GALLAGHER,AND E. BALECH. 1994. Biogeography of toxic dinoflagellates

    in the genus Alexandriumfrom the northeastern United Statesand Canada. Marine Biology 120:467478.

    ANDERSON, D. M., D. M. KULIS, J. J. SULLIVAN, AND S. HALL. 1990.Toxin composition variations in one isolate of the dinoflagel-late Alexandrium fundyense. Toxicon 28:885893.

    ASHWORTH, W. 1986. The Late, Great LakesAn EnvironmentalHistory. Alfred A. Knopf, Inc., New York.

    BERG, G. M., P. M. GLIBERT, M. W. LOMAS, AND M. A. BURFORD.1997. Organic nitrogen uptake and growth by the chryso-phyte Aureococcus anophagefferens during a brown tide event.Marine Biology 129:377387.

    BODEANU, N. 1993. Microalgal blooms in the Romanian area ofthe Black Sea and contemporary eutrophication conditions,p. 203209. In T. J. Smayda and Y. Shimizu (eds.), Toxic Phy-toplankton Blooms in the Sea. Proceedings of the Fifth In-ternational Conference on Toxic Marine Phytoplankton. El-sevier, Amsterdam, The Netherlands.

    BODEANU, N. AND G. RUTA. 1998. Development of the plankton-ic algae in the Romanian Black Sea sector in 19811996, p.188191. In B. Reguera, J. Blanco, M. L. Fernandez, and T.

    Wyatt (eds.), Harmful Algae. Xunta de Galicia and Intergov-ernmental Oceanographic Commission of United Nations Ed-ucational, Scientific and Cultural Organization, Paris, France.

    BOYER, G. L., J. J. SULLIVAN, R. J. ANDERSEN, P. J. HARRISON, ANDF. J. R. TAYLOR. 1987. Effects of nutrient limitation on toxinproduction and composition in the marine dinoflagellate Pro-togonyaulax tamarensis. Marine Biology 96:123128.

    BOYNTON, W. R., J. H. GARBER, R. SUMMERS, AND W. M. KEMP.1995. Inputs, transformations, and transport of nitrogen andphosphorus in Chesapeake Bay and selected tributaries. Es-tuaries18:285314.

    BOYNTON, W. R., W. M. KEMP, AND C. W. KEEFE. 1982. A com-parative analysis of nutrients and other factors influencingestuarine phytoplankton production, p. 6990. In V. S. Ken-

    nedy (ed.), Estuarine Comparisons. Academic Press, NewYork.

    BURFORD, M. A. 1997. Phytoplankton dynamics in shrimp ponds.Aquaculture Research 28:351360.

    BURFORD, M. A. AND P. M. GLIBERT. 1999. Short-term nitrogenuptake and regeneration in early and late growth phaseshrimp ponds. Aquaculture Research 30:215227.

    BURKHOLDER, J. M. 1998. Implications of harmful microalgaeand heterotrophic dinoflagellates in management of sustain-able marine fisheries. Ecological Applications 8:S37S62.

    BURKHOLDER, J. M. 2000. Eutrophication and oligotrophication,p. 649670. In S. Levin (ed.), Encyclopedia of Biodiversity,

    Volume 2. Academic Press, New York.BURKHOLDER, J. M. AND H. B. GLASGOW. 1995. Interactions of a

    toxic estuarine dinoflagellate with microbial predators andprey. Archiv fur Protistenkunde 145:177188.

    BURKHOLDER, J. M. AND H. B. GLASGOW. 1997. The ichthyotoxic

    dinoflagellate Pfiesteria piscicida: Behavior, impacts and envi-ronmental controls. Limnology and Oceanography 42:10521075.

    BURKHOLDER, J. M. AND H. B. GLASGOW. 2001. History of toxicPfiesteriain North Carolina estuaries from 1991 to the present.BioScience51:827841.

    BURKHOLDER, J. M., H. B. GLASGOW, AND N. J. DEAMER-MELIA.2001a. Overview and present status of the toxic Pfiesteriacom-plex. Phycologia 40:186214.

    BURKHOLDER, J . M., H. B. GLASGOW, N . J . DEAMER-MELIA, J.SPRINGER, M . W. PARROW, C . ZHENG, AND P. CANCELLIERI.

  • 8/3/2019 Harmful Algal Blooms and Eutrophica