THE LIGHT REQUIREMENTS FOR GROWTH AND PHOTOSYNTHESIS IN SEAGRASSES WITH EMPHASIS ON TEXAS ESTUARIES: A LITERATURE SURVEY J.E. Kaldy and K.H. Dunton University of Texas at Austin Marine Science Institute Port Aransas, Texas 78373 Submitted to: United States Environmental Protection Agency Region 6 1445 Ross Avenue, Suite 1200 Dallas, Tx 75202-2733 31 August 1993 THE LIGHT REQUIREMENTS FOR GROWTH AND PHOTOSYNTHESIS IN SEAGRASSES WITH EMPHASIS ON TEXAS ESTUARIES: A LITERATURE SURVEY J.E. Kaldy and K.H. Dunton University of Texas at Austin Marine Science Institute Port Aransas, Texas 78373 Submitted to: United States Environmental Protection Agency Region 6 1445 Ross Avenue, Suite 1200 Dallas, Tx 75202-2733 31 July 1993 Table of Contents Acknowledgements 3 Executive Summary 4 Chapter 1. INTRODUCTION 9 The Texas flora 9 seagrass 21 Importance of seagrasses LIGHT AND WATER TRANSPARENCY 34 Chapter 2. Environmental influences on water clarity 34 light in the aquatic environment 34 Natural influences 40 on water clarity on 43 Anthropogenic influences water quality and water clarity THE CHESAPEAKE BAY APPROACH 50 Chapter 3. 50 Approaches to assessing habitat requirements of seagrasses SEAGRASS PHOTOSYNTHETIC PHYSIOLOGY 54 Chapter 4. Computer modelling 57 Effects of in situ light reduction 58 63 Experimental changes in daily light period 68 Laboratory studies of reduced light on seagrasses 70 .Photosynthesis irradiance curves 85 Seasonal changes in seagrasses SYNTHESIS AND RECOMMENDATIONS FOR Chapter 5. FURTHER RESEARCH 97 literature Cited 101 Acknowledgements Numerous individuals contributed successful of this greatly to the completion literaturereview. WeareverygratefulfortheconstructivecommentsprovidedbySteffanie Barnett,Philip Crocker, KenTeague andCarlYoungofthe U.S.Environmental Protection Agency (Region 6) and Chris Onuf (U.S. Fish and Wildlife Service) and Warren Pulich Parks and (Texas Wildlife Department). UTMSI graduate students in marine botany, Andrew Sharon Herzka and Lee including Czerny, Kun-seop provided many helpful comments as our unofficial editors; Patty Baker and Kim Jackson were instrumental in the We also typing and preparation of the initial drafts that eventually led to the final report. of the hard-to-find thank the UTMSI library staff for their expedient acquisition for many books and papers referenced in this review. The cover sketch was prepared by Nancy Buskey. This work was supported by the U.S. EPA Region 6 Grant No. X-996025-01-0 to Ken Dunton. Correct citation of this report is: Kaldy, J.E. and K.H. Dunton. 1993. The light requirements for growth and photosynthesis in seagrasses with emphasis on Texas estuaries: a literature survey. U.S. EPA Region 6, Dallas, Texas. Executive Summary Seagrasses have been repeatedly demonstrated to be highly valuable components of coastal systems. They have high rates of primary productivity and support a diverse assemblage ofconsumers as well as rapidly cycling ecologically important elements such as carbon,nitrogen,phosphorusandsulphur. Thehighprimaryproductivityofseagrasssystems supports many commercially and recreationally important species. In addition, they also act and reduce to modify the deposition of sediments, remove nutrients, attenuatewave energy currents. Ingeneral,seagrassecosystemsareacornerstoneofhealthy,productivebaysand estuaries. the last 20 communities the world have During years, seagrass throughout experienced decreased productivity and distribution. These declines have often been as a attributed to decreased water transparency result of turbidity or shading by epiphytic is often an indication of nutrient enrichment caused algae. Epiphytic shading by anthropogenic inputs. Although both epiphytes and turbidity occur as natural phenomena, human activities can exacerbate existing natural conditions with adverse effects on seagrass communities. The objectives of this study were (a) to review the existing literature and data available on the effect of natural and anthropogenic factors the underwater light on distribution and environment; (b) to examine the relationship between light and seagrass productivity; and (c) to make recommendations on how to habitats in Texas protect seagrass bays and estuaries. To meet these goals, we have examined the available literature, Table 1. list of all submerged aquatic plant species (in italics) cited in the text. Species listed by family and include the author who first described each species (not shown in are italics). Seagrasses Cymodoceaceae Amphibolis antarctica (Labill.) Sonder ex Aschers. Cymodocea nodosa (Ucria) Aschers. Halodule wrightii Aschers. Syringodium filiforme Kutzing Hydrocharitaceae Enhalus acoroides (Linnaeus f.) Royle HalophUa decipiem Ostenfeld Halophila engelmannii Aschers. HalophUa johnsonii Eisman HalophUa stipulaceae (Forsk.) Aschers. Thalassia testudinum Banks ex Konig Posidoniaceae Posidonia angustifolia Cambridge et Kuo Posidonia australis Hook. F. Posidonia oceanica (L.) Delile Posidonia sinuosa Cambridge et Kuo Zosteraceae Heterozostera tasmanica (Martens ex Aschers.) den Hartog Phyllospadix scouleri Hooker Phyllospadix torreyi S. Watson Zostera angustifolia (Hornem.) Reichb. Zostera marina L. Zostera muelleri Irmisch ex Aschers. Freshwater and Estuarine Plants Hydrocharitaceae Elodea canadensis Michx. Vallisneria americana Michx. Potamogetonaceae Potamogeton pectinatus L. Potamogeton perfoliatus L. Ruppia maritima L. emphasizing the physiological response of seagrasses to light and temperature. By using data and observations collected on a variety of species (Table 1) from around the world we A may be better able to define the light requirements of Texas seagrasses. knowledge of the minimum annual light requirements for seagrass growth is necessary to maintain the current distribution of Texas This information will also be species. required in the development of a management plan that permits the expansion and establishment of new habitat. seagrass Five species of seagrasses from three families comprise the submerged vascular flora of the Texas some acres bays and estuaries, encompassing 209,738 (Onuf 1993). Most of these species have tropical or sub-tropical affinities and are near the northern end of their distribution in Texas. One species, Ruppia maritima is considered , cosmopolitan and occurs as far north as New Hampshire, while Halodule wrightii occurs as far north as North Carolina. Most seagrasses are perennial; however, Ruppia has been Texas estuaries. reported as both an annual and a perennial in some Although flowering in Thalassia testidinum has been documented only once in Texas (Phillips et aL, 1981), all five known to flower and set seed. are seagrass species Lunar tides seldom exceed 15 cm in Texas bays and estuaries, and therefore are heavily influenced by local meteorology. Meterological events (particularly winds and storms) also control the turbidity of the bays along the Texas coast. However, anthropogenic factors can also influence turbidity. Dredging activities can increase light attenuation and prevent the plants from receiving their daily minimum light requirements for growth and survival. In Texas, increased turbidity as result of dredging and a construction appears to be the largest threat to seagrasses. Presently, other anthropogenic influences such as eutrophication, which results from increased nutrient loading, occurs only in a few localized areas (i.e. portions of the Galveston Bay System and Copano Bay). increases, does the threat of as zone so However, the human population of Texas coastal eutrophication. Severalrecentpublications haveproposedanovelmethodforassessingwaterquality in estuarine systems (Batuik et al, 1992; Dennison et al, 1993). The basic premise of the in areas technique is that plants will not where their habitat requirements (i.e. light, grow nutrients etc.) are not adequatly met. Thus, if the minimum habitat requirements are be known, the parts of the system that are not meeting the minimum requirements can inferredusingmappingtechniques. Someaspectsofthistechniquemaybeappropriatefor are not. useinTexassystems,butmany Thisisaresultofthemanagementhistoryofeach estuarine system; Chesapeake Bay is highly impacted, and the goals are to restore and conserve the remaining resources. Texas systems, with some exceptions, are relatively pristine and preservation of the existing seagrass habitats is the major goal. In general, lower irradiance results in reduced density and biomass of seagrasses. Laboratory and field studies show that seagrasses may also respond to a drop in irradiance by altering their morphology. The minimum light requirements of seagrasses can be quantified through measurement of the daily minimum number of hours of saturating irradiance needed to meet their respiratory demands. The H value for some ml seagrasses (i.e. Zostera marina) may approach 6 h to maintain a positive carbon balance; been determined. however, the requirements for most Texas species have not In Texas, Halodule wrightii has recently been shown to have an H of 3to 5 h (Dunton, submitted). sat A review of the literature suggests that most seagrass species, including those along the Texas coast, require 10 to 20% of surface irradiance as an absolute minimum during most of the year. Although light is a critical factor controlling seagrass growth and distribution, A temperature is also important. light level that is photosynthetically saturating (Ik) at low temperatures may be the compensating irradiance (I) at a higher temperatures. As a result c of seasonal variation in temperature, most seagrasses exhibit seasonal patterns with respect to productivity, P vs. I parameters and organic composition. The temporal dynamics of are such that the winter to seagrasses appears beaperiodoflowphysiologicalactivity,while the extent of photosynthetic activity in spring, summer and fall determine the long-term success of a species. We recommend that human disturbance be kept to a minimum during critical periods of seagrass growth, which is greatest during spring and early summer. CHAPTER 1; INTRODUCTION The Texas seagrass flora Seagrasses are flowering plants which re-invaded the marine environment. The - date from the Cretaceous about 140 million earliest fossils of seagrasses years ago (den Hartog, 1970). Unlike the algae, which are in the Kingdom Protista, seagrasses are vascular plants that have xylem and phloem elements for the transport ofwater and photosynthate (Pedersen and Sand-Jensen 1993, Barnabas and Araott 1987, Barnabas 1989,1991). These plants also differ from algae in the degree of cellular differentiation and organization; have true roots, stems, and leaves. these seagrasses Algae do not possess organs. As marine angiosperms (flowering plants), seagrasses complete their life-history underwater. They reproduce sexually by flowering and vegetatively by rhizome branching. Rhizome formed the initiation of become branches, by axillary meristems, may via physiologically independent and create new plants. In plants, sexual reproduction occurs flowering and pollen exchange. To exchange pollen in the aquatic environment, plants have evolved three strategies (Fig. 1). These include (Cox, 1988) the transport of pollen above the water surface (category I), pollen transported on the water surface (category II), and pollentransportedbeneaththewatersurface(categoryIII). CategoryIpollentransporthas beenreportedforonlyone speciesofseagrass(Enhalusacoroides);however,itiscommon in freshwater submerged aquatics. Most seagrasses exhibit either category II or 111 pollen transfer. Seed germination occurs underwater and is influenced by both temperature and Figure 1. Schematic diagrams of category 2 (top panel) and 3 (bottom panel) hydrophilly. Category2hydrophillyexemplifiedbyHalodulepinifolia(toppanel). A.Erectionofanother at low tide; B. Another dehiscence and assembly of search vehicles; C. Search vehicles floating onwatersurface withinsert showing darkfieldsilhouette; D.Pollinationbycollision of search vehicles with filamentous stigmas; E. Close-up of filiforme pollen. Category 3 hydrophilly exemplified by Thalassia testudinum (bottom panel). A. Male flowers; B. Mucilage string containing pollen; C. Underwater dispersal of mucilage strand; D. Female flowers with rigid stigmas; E. Pollination by collision. From Cox 1988. 10 The details of seed and germling anatomy have been described for Zostera marina salinity. (Taylor 1957a,b) and Thalassia testudinum (Orpurt and Boral 1964). In the early development of the seedling, seagrasses exhibit a single cotyledon ("seed leaf'). Thus seagrasses aremonocotswhichhavebeenplacedintheclassliliopsida, subclassAlismatidae (Cronquist 1981). CJ.G. Petersen and colleagues were the first researchers to realize the importance ofcommunities to coastal ecosystems (Petersen 1891,1918; as seagrass cited in den Hartog 1980). They suggested that seagrass production (8 million dry tons/year) was the basis of the food web for all marine fauna (Petersen and Boysen-Jensen 1911, Petersen 1913,1915, 1918, Boysen-Jensen 1914; as cited in Rasmussen 1977). However, they underestimated the importanceofphytoplankton, asreflectedintheabsenceoffinfisheryfailurewiththeonset on of the wasting disease of the 1930’s (Rasmussen 1977). Much of the other early work was seagrasses related to taxonomy (Ascherson 1907, Hutchinson 1934, Markgraf 1936; as citedindenHartog 1970),whilemuchofthephysiologicalandecologicalresearchhasbeen conducted since the late 1960’5. There are that between the Arctic and Antarctic occur about 48 species ofseagrasses Circles (Phillips and Menez 1988). Many of these species have a circumglobal distribution (e.g.Zosteramarina whichoccursonbothsidesoftheAtlanticandthePacific;denHartog , 1970). Almost all seagrass species display continuous areas of distribution (Phillips and Menez 1988). However, some species have disjunct distributions, which may be due to the movement of the continents during the geological epochs (den Hartog 1970). Because of the time scales inferential data and fossil foramanifera involved, data (i.e. geological associated with seagrasses) have been used to account for the present day distribution of Like seagrasses (Phillips and Menez, 1988). many other organisms, the areas of highest are tropical and subtropical environments (den Hartog 1970). There are species diversity that are considered characteristic of tropical areas and five that are seven genera genera considered temperate in character (Table 2). Although the Pacific coast of South America and the Caribbean have relatively high species diversity, they are notas diverse at the Indo­ most areas West Pacific. The seagrass flora of of the world is well documented, except for theAtlanticCoastofSouthAmerica,whichremainsrelativelyunexplored(denHartog 1970, Phillips and Menez 1988). Strictly speaking, there are four seagrasses (Thalassia testudinum , Syringodium Halodule one filiforme, wrightii and Halophila engelmanni) and euryhaline aquatic plant, Ruppia maritima that occur along the Texas coast. Because R. maritima is found in , freshwater, neither den Hartog (1970) nor Phillips and Menez (1988) consider R maritima a seagrass. However, Ruppia does grow and complete its life history in many of the of this review hypersaline bays and estuaries of Texas (Dunton 1990) and for the purpose will include Ruppia with the we seagrasses (Table 3; Figure 2). With the exception of the Texas Ruppia, seagrasses have tropical or subtropical affinities. Ruppia is cosmopolitan and occurs as far North as New Hampshire (J. Kaldy, pers. obs.), while Halodule wrightii occurs as far north as North Carolina. The seagrasses also tend to be perennial (Table 4); annual although, Ruppia in Texas has been reported as a perennial (Pulich 1985) and as a Thalassia (Dunton 1990). Texas seagrasses, except exhibit category II pollenation; it is , unclear whether Ruppia exhibits category II or 111 pollen transfer. Table 2. The affinities of the major seagrass species. Some of the species with tropical affinities overlap into subtropical or warm areas. From den Hartog 1970. Tropical Genera Temperate Genera Enhalus Halodule Thalassia Heterozostera Halophila Phyllospadix Cymodocea Posidonia Zostera Syringodium Thalassodendron Table 3. A listing of Texas seagrass species. Family and Species Common Name Cymodoceaceae Halodule wrightii Aschers. Shoal grass Manatee Syringodium filiforme Kiitzing grass Hydrocharitaceae Halophila engelmanni Aschers. Clover grass Thalassia testudinum Banks ex Konig Turtle grass Potamogetonaceae Ruppia maritima L. Widgeon grass Figure 2. Schematic drawings of the seagrass flora of the Texas coast. Modified from Zieman 1989. Table 4. Some of the characteristics of Texas seagrasses. Common name Pollen category Biogeographic affinity Longevity Clover grass II tropical perennial Manatee II grass tropical perennial Shoal grass II subtrop-trop. perennial Turtle grass m tropical perennial Widgeon grass ii, in cosmopolitan annual/perennial Mostseagrasses, includingTexas species,growinprotectedbays andestuaries,where andrhizomestructureoftheplantspermitsthepenetrationandcolonization ofsoft the root bottom substrate (i.e. sand and mud). However, there is at least one that inhabits the genus rocky intertidal along the high energy coasts of the Pacific (i.e. Phyllospadix). The species hair ofPhyllospadix have special adaptations, such as extensive hypodermal fibers and root rhizomes and small lacunae to colonize the crevices of rocks development, thickened (Cooper and Mcßoy, 1988). With the exception of Phyllospadixmost seagrasses require , sediment depths between 5 and 25 cm for adequate anchoring (Zieman 1972). Sediments are classified as either terrigenous (i.e., derived from terrestrial sources) or carbonate. Carbonate sediments are usually biogenic in origin (Scoffin 1970). Seagrasses also obtain atleastpartoftheirinorganic nutrientsfromthesediments(Thursby andHarlin 1982;Short and Mcßoy 1984). Seagrasses colonize a relatively broad range of depths (Figure 3), depending upon available light and substrate type as well as the physiological requirements of the species. IntherelativelyclearwatersoftheMediterraneanSea,Posidonia hasbeenreportedto grow below about 1.5 at depths of 50 m (Gessner 1961). However, Zostera marine does not grow mintheturbidwatersofSanFranciscoBay,(Zimmermanetal.,1991). Meanwhile,Ruppia maritima along the Texas Coast has been reported growing at 0.3 m above mean low water (C. Belaire, 1993, pers. comm). Thus, the depth distribution of a species is variable and several factors. depends upon Although neither Posidonia nor Zostera occur along the Texas coast, they serve as examples to show the extreme range of depths that these plants are able to colonize. Figure 3, The reported depth limits of 31 marine angiosperm species. Bars represent the of values encountered, while the solid value for each range square represents the average species. Number in right column indiciates the number of estimates. From Duarte 1991. Several authors have examined the depth distribution of Thalassia testudinum within the Gulf of Mexico. Buesa (1974) examined the population and biological parameters of Thalassia testudinum on the Cuban shelf. He found that Thalassia had a biomass of m'nr approximately 200 g 2 to a depth of 1 m, while the lowest biomass, less than 50 g 2, occurred at 14 m depth (Figure 4). The depth distribution of other species varies considerably. For example, Halophila decipiens is found to a depth of 25 m (Figure 4). Buesa (1974) also suggested that Thalassia would not grow in areas where irradiance at the seabed was less than 25% of surface irradiance. Vicente and Rivera (1982) examined the depth limits of Thalassia growing in Puerto Rico and found a statistically significant positive mean Secchi depth and the lower limits of Thalassia (see chapter 2). correlation between Thus, they suggest that where the water is clearer (deeper secchi depth) the plants will colonize at lower depths. However, they also suggest that herbivory may limit the depth distribution of Thalassia. Dawes and Tomasko (1988) investigated the depth distribution of Thalassia in Florida. They found that plants collected from the deep edge of the bed had lowershoot density, greaterleafarea,andgreaterabove:belowgroundbiomass than plants collectedfromtheshallowareas,suggestingthatdeepplantsweremorelight stressedthan shallow plants. It is generally accepted that light availability is the factor that controls the depth distribution of seagrasses. For example, Buesa (1974) concluded that light energy and temperature were the factors that control the depth distribution of Thalassia. Vincente and Rivera (1982) also suggested that Thalassia was limited to areas where irradiance at the plant depth was adequate. Iverson and Bittaker (1986) proposed that the depth Figure 4. The depth distribution of four seagrass species on the northwest coast of Cuba. 1 = Thalassia testudinum 2 = 3 = 4 = H. Syringodium filiformeHalophila decipiens , ,, From Buesa 1974. engelmanni. distribution of Thalassia was light limited. Orth and Moore (1988) suggested that species zonation was a physiological response to the interaction of light and temperature. Importance of seagrasses We are interested in defining the light requirements of seagrasses because they are an important component of estuarine and coastal ecosystems (Zieman 1982; Phillips 1984; Thayer et al., 1984; Zieman and Zieman 1989). There is a large body of literature documenting the beneficial qualities of seagrass beds. The three-dimensional habitat that beds create is important to species of fish and shellfish including seagrass many commercially important species (Figure 5). There were two studies during the 1970’s that beds. documented the importance of seagrass Thayer et aL (1975) found that eelgrass beds support numerous types of macrofauna and epifauna, which may consume more than half ofthe net productionoftheeelgrass-plankton-algalsystem. Rassmussen(1977)showedthat the species composition and abundance of invertebrate organisms changed with the loss of eelgrass (Zostera marina) due to the 1930’s outbreak of the wasting disease. During the 1980’s several other investigators documented the importanceof seagrass habitat. Orth et al (1984) compared unvegetated areas with seagrass meadows and found that seagrass beds contained a dense and rich assemblage of vertebrates and invertebrates. Species abundance was positively correlated with two aspects of plant morphology: 1) the root-rhizome mat, and 2) the plant canopy. The increased species diversity and abundance suggests that the three-dimensional habitat was beneficial to the on Figure 5. Proposed food web for a seagrass bed, based dominant organisms of Indian River Lagoon, Fla. The grazing amphipod (Cymadusa comptd) is about to be preyed upon by the shrimp (Palamonetes intermedins) which is to be preyed upon by the fish (Bairdiella chrysoura ). From Virnstein 1987. secondary productivity of the system. Vimstein and Curran (1986) studied the colonization ofartificial habitat.Theyfoundrapidcolonizationofsubstratewithmaximum seagrass new diversity and abundance within 4-8 days, suggesting that new seagrass substrate may enhance Eckman secondary productivity within an estuary. (1987) found that eelgrass (Zostera marina) changed the hydrodynamics ofwatercurrents, facilitating the recruitmentofscallops (Argopecten irradians) and the commonjingle (Anomia simplex). Eckman also suggested that the hydrodynamic influence of seagrass was more important than predation in determining the abundance of recruits to the system. Short (1988) found that scallops actively migrated beds. into transplanted seagrass During the 1990’s there have been several more investigations ofthe importance of seagrass habitat. Pohle et aL (1991) found that juvenile scallops actively attached themselves to eelgrass blades above the substrate as a refuge from predation. Sogard and Able (1991) demonstrated that vegetated substrate (Zostera or Ulva) was superior in quality (based on fish and decapod densities) to adjacent unvegetated substrate. Sites with Zostera as the dominant plant had higher densities of most fish species than areas dominated by Ulva; however, they concluded that Ulva was important in areas lacking seagrass Hoven cover. (1992) suggested that eelgrass (Zostera marina) meadows are of considerable importance as sites for the settlement of blue mussel (Mytilus edulis) larvae. of these Although many studieswereconcerned withcommerciallyimportantspecies,otherspeciesalsobenefitfrom habitat. seagrass In general, healthy seagrass systems help to create and foster secondary production. Seagrassesact as asubstrate forthe growthofalgal epiphytes which contribute large 13 amounts of fixed organic carbon to both the and detrital food webs. C grazer Using Thalassia > where Rhizophora prop Thalassia with heavy epiphytes > Laurencia > Polysiphonia. Mangroves (Rhizophora) and seagrass ( Thalassia ) were most efficient at reducing current flow. The ability of Thalassia to reduce current velocity was directly proportional to the density of seagrass blades; current velocities > 150 cm s' 1 were required to erode sediments from dense Thalassia beds (Scoffin 1970). To experimentally assess the impact of seagrass beds on currents, Fonseca et al (1982) used a salt water flume. They foundapredictablereductionintheapparentcurrentvelocitybyseagrass andsuggestedthat current reduction properties may vary between species and sites, due to morphological and Ward et hydrographical variation (Fonseca et al, 1982). al (1984), working in Chesapeake Bay found that attenuated wave As seagrasses energy and inhibited sediment resuspension. a result, sedimentation rates were substantially higher in seagrass communities than in unvegetated areas. Fonseca and Cahalan (1992) evaluated wave attenuation by four seagrass species andfoundthatbroad,shallowseagrass meadowssubstantiallyattenuateup of bed. to 40% of wave energy per meter seagrass Recent research has also shown that communities act as seagrass a "biological scrubber." Short and Short (1984) found that seagrasses, growing in mesocosms, rapidly filtered out sediments that were added to the water column, resulting in increased light penetration. Additionally, they found that seagrasses rapidly removed nutrients added to the filtered water column (Figure 8). Thus, the seagrasses out both sediments and nutrients. In general, research has shown that seagrasses are a cornerstone to the health and productivity of estuarine communities (Zieman 1982; Phillips 1984; Thayer et at, 1984; Zieman and Zieman 1989). They provide habitat and food to a wide variety of organisms. Seagrasses also physically alter the environment they inhabit, by influencing water currents and sedimentation processes. However, the long-term future of these communities is in jeopardy. world wide decline in For Recently, there has been a seagrass habitat (Figure 9). example,OrthandMoore(1984)showthatbeforethe 1960’5,submergedaquaticvegetation (SAV) was a widespread feature of the Chesapeake Bay System. However, Percent of surface light extinction and coefficients in a Figure Ba. Halodule wrightii tank, an a Syringodium filiforme tank and unvegetated tank following addition of Indian River From Short and sand-silt sediment (a), clay sediment (b), and organic-silt sediment (c). Short 1984. Figure Bb. Nutrient removal from the water column by the seagrass community. Changes in phosphate (a) and ammonium (b) concentration for Halodule and a nonseagrass control tank over time. From Short and Short 1984. 30 time in Australia. From Cambridge and McComb Figure 9. Changes in seagrass cover over 1984. since the 1960’s there has been a dramatic loss of SAY communities. Livingston (1987) have been virtually eliminated from Pensacola and Tampa Bays. reported that seagrasses Eleuterius (1989) has documented the loss of about 70% of the seagrass habitat from 1989. Mississippi Sound between 1969 and During the 1930’s up to 90% of the eelgrass populations occurring along the Atlantic Coast were lost (Costa 1988). The decline of eelgrass was due to infection by the "wasting disease" caused by a marine slime-mold Labyrinthula zosterae (Muehlstein et al, 1991). The wasting disease was a natural caused Disease catastrophe, by the parasitic organism. still plays a role in controlling declines in seagrass distribution (Short et al., 1991); however, the more recent seagrass habitat are due primarily to anthropogenic influences. Aerial photography and ground truthing have been used to document changes in the distribution and biomass of seagrasses. Merkord (1978) documented changes in seagrass distribution of the Laguna Madre system between 1965 and 1976. In the Upper Laguna Madre, Halophila and Halodule cover increased, while Ruppia cover decreased. In Lower Laguna Madre, Syringodium cover increased, displacing Halodule in some areas while Thalassiaexpandeditsrangenorthward. Additionally,significantportionsofPortIsabelBay have become bare devoid of cover. that the areas, seagrass Merkord (1978) suggested changes were a result of decreased salinity (over the long term) and increased turbidity. Costa (1988) documented changes in the abundance of eelgrass related to anthropogenic and natural disturbance using aerial photographs, charts, writtenreports, local residents and sediment He suggests that the wasting disease took a massive toll on eelgrass cores. populations, eliminating about 99% of the eelgrass in Buzzards Bay. Pulich and White (1991) have documented the decline of seagrass habitat in the Galveston For Bay System. They attribute various processes to the loss of specific seagrass habitat. example, the loss of Ruppia maritima was related to Hurricane Carla and a rise in relative level due to subsidence. However, in the lower bay both Halodule wrightii and Ruppia sea maritima disappeared between the 1950’s and the 1980’s. These declines may be due to increased human activities such as urban development, wastewater discharges, chemical spills and dredging activities. Quammen and Onuf (1993) documented increase of 140 an km2 of bare bottom in Laguna Madre, Texas. They suspect that light reduction from maintenance dredging has caused the loss of cover. Thus, historical data bases, seagrass nautical charts, and aerial photography have proven useful in documenting changes in the distribution of seagrasses. Presently, declining water quality (both transparency and nutrients) is adversely as a affecting seagrasses. In general, reduced light availability result of anthropogenic influence is the greatest threat to seagrasses worldwide (Merkord 1978; Cambridge and McComb, 1984; Costa 1988; Giesen et al, 1990; Pulich and White 1991; Short et aL, 1991; Quammen and Onuf 1993). Thus, before seagrass can be expected to recolonize (either with or without human intervention), water quality problems must be addressed. CHAPTER 2; LIGHT ANDWATER TRANSPARENCY Environmental influences on water clarity Estuarine biologists and managers often refer to the concepts of "water quality" and "water somewhat This is unfortunate, because there is a clarity" interchangeably. fundamental difference between the two expressions. Water quality refers to the chemical and physical parameters (i.e. nutrient concentrations, dissolved oxygen, salinity, temperature, etc.) that characterize a parcel of water with respect to the effect of these parameters on the the health of aquatic organisms, in this case plants. Water clarity or transparency, on other hand, is a specific character of water quality. It is defined by the amount of light transmitted through a body of water. Decreased water quality (e.g. increased nutrient can concentration) stimulate phytoplankton blooms which reduce light transmittance (e.g. water clarity). Thus, water clarity is related to and influenced by water quality, but these will terms are not the same. Throughout this document we use the terms water quality and consistent with the above definitions. water clarity in a manner Light in the aquatic environment Photosynthetically active radiation (PAR) is that portion of the electromagnetic PAR extends from about 350 nm to about spectrum utilized by plants for photosynthesis. 700 nm wavelength and is roughly equivalent to the range of wavelengths to which the human eye is sensitive (i.e. visible light). Seagrasses require light energy for the process ofphotosynthesis. However,thereareseveralfactorsthatinfluencetheamountoflightthey The receive; three of these major factors are: (1) albedo; (2) scattering and (3) absorption. portionoflightreflectedbackinto isreferredtoasthealbedoofthatsurface(Figure space 10). The albedo of the Atlantic Ocean at 30° North latitude is approximately 0.068 (Payne, surface at this 1972). Thus, approximately6.8%ofthelight energyimpingingontheocean latitude is reflected back to space and is unavailable to marine plants, light is also scattered within the water column by suspended particles (Figure 11). In addition to albedo as it and scattering, light is selectively absorbed passes through the water column (Figure 12). used to measure the light field underwater because Spherical quantum sensors are they measureboth downwelling and scattered light. Flat cosine sensors underestimate light availabilitybecausetheyaccountonlyfordownwellingirradiance. Becauseofscatteringand selective modelled with wavelength attenuation, light penetration through water is an exponential decay function (Kirk 1983): (b) I* I»e­ = where \is the irradiance at depth z, I is the incident irradiance at the water-atmosphere Q interface, and k is the light attenuation coeffecient. Simulation models and field data have been used in the investigation of light column of estuaries. simulation model penetration through the water Hogan (1983) used a nm in of light attenuation and found that maximum transmissivity occurred at about 465 clear water and at 550 nm in turbid water (Figure 12b). He concluded that the shift was Figure 10. Schematic diagram showing the different origins of light received from a remote sensor above the water. The reflection of the direct solar beam at the surface is termed albedo. From Kirk 1983. 11. Figure light availability underwater is determined by attenuating processes. Light attenuationresultsfromtheabsorption andscatteringoflightbyparticlesinthewater(i.e., suspended solids, phytoplankton, etc.) as well as the absorption of light by the water itself. From Dennison et al 1993. ., 37 Figure 12a. The spectral range of underwater irradiance decreases with increasing depth in clear oceanic water. At 10 m about 70% of the spectrum is fairly broad while at 90 m the quanta is in the 450-500 run band. From Saffo 1987. Figure12b. Sectralchangesofunderwaterirradiancewithincreasingwaterturbidityat10 in m depth. Moderate turbidity (3) results in maximal transmittance green portion of spectrum while under conditions of maximal turbidity (type 9) transmitted light is mostly yellow. From Saffo 1987. 38 due to suspended matter in the water, which absorbed and scattered shorter wavelengths more The ultimate result was that the maximum transmittance of than longer wavelengths. occurred in the in turbid coastal while total light longer wavelengths environments, transmittance decreased. Pierce et al (1986) investigated how light transmittance respondedtochangesintheturbidityoftheRhodeRiver Estuary. lighttransmittance,both spectral quality and intensity, varied with changes in the amount of dissolved and suspended High concentrations of suspended solids and dissolved materials were correlated matter. with increased attenuation in the upper water column; attenuation varied with wavelength depending on the materials present. Regression analysis indicated that the concentration ofchlorophyllaandc,aswellasmineralmatter,accountedformostofthevariation. These investigations suggest that light transmittance in turbid estuarine waters is more complex than clear oceanic water. Recently, many authors have estimated attenuation coefficients from Secchi depths as a means of calculating light availability (Chambers and Kalff 1985; Duarte 1991; Batuik et al, 1992; Dennison et al, 1993 and others). Preisendorfer (1986) examined the theory and mathematics behind the use of the Secchi disk. He suggested that use of the Secchi disk to evaluate the attenuation coefficient, k, obliterates and abuses the primary function of the Secchi disk. He concluded that transparency measurements made with a Secchi disk do not yield valid information on the availability of light at depth. Therefore, long term in situ electronic light measurements are required to estimate underwater light availability. However, other investigators (Megard and Berman 1989) suggest that Secchi depth (transparency measurements) is proportional to the attenuation coeffecient for the most conducted penetrating waveband. However, the work of Megard and Berman (1989) was in oceanic waters and was not subject to the same scattering problems endemic to nearshore and estuarine systems. Historically the Secchi depth has been taken as the 18% light level; however, new calculations suggest that Secchi depth is the 22% light level (Megard and Berman Due 1989). to the ongoing literature debate, the applicability of Secchi depth environments As a researchers measurements to estuarine is questionable. result, are starting to measure in situ light availability directly using photoelectric cells, such as spherical quantum sensors (Tomasko and Dunton, 1991). A long term (four-year) data set now exists for Upper Laguna Madre, Texas (Dunton, submitted). Natural influences on water clarity As mentioned previously, three factors control how much light reaches submerged Albedo varies with latitude and season due vegetation: albedo, scattering and absorption. to the declination of the earth with respect to the sun. However, scattering and absorption of light within the water column are affected by a number of natural and anthropogenic influences. tides and floods increase Currents, may the suspended solids, reducing the amount of light reaching the plants. In a rather extreme case, the onshore movement of coastal sand smothered a seagrass bed in Australia (Kirkman, 1978). Wind events have also waves can cause beencitedasinfluencinglight;wind theresuspensionofsedimentsresulting in high light attenuation (Ward et aL, 1984). Severe wind events, in the form of hurricanes, havebeencitedashavingadverse effects onseagrasses and lightinFlorida(Zieman 1975a) and in Mississippi Sound (Eleuterius 1989). In a recent manuscript Onuf (1993 in review). had to incorporate wind speed and direction into his statistical analysis to accurately model the propagation of dredging effects in Laguna Madre, Texas. and Meteorological forcingintheformofwindwaves, large-scalewind-driven gyres flushing due to frontal passage have been shown to strongly influence the hydrography of Gulf estuaries (Ward and Armstrong 1980). The influence of meteorological forcing is a consequence of large surface area to volume ratios as well as the intensity and variability Wind driven waves are a of meteorological events (Ward 1979). result of meteorological forcing; light to moderate winds over long fetches allow the development of intense surface waves. The mixing action of these wind waves results in waters that are usually vertically homogenous, except in the deeper dredged channels (Ward 1979; Ward and Armstrong 1980). The seasonality of winds may also play a role in controlling light attenuation; for example,Riceetal. (1983)found thatsediment resuspension occursyear-roundbutthatit may be most active during winter storms. The seasonality of winds along the Texas coast is governed by the intensity of the Bermuda High (Ward and Armstrong 1980). As a result, light transmittance during certain times of the be affected more by sediment year may suspensionthanduringotherpartsoftheyear(RiceetaL, 1983;Wardetaly1984). Water clarity in bays and estuaries along the Texas coast is mostly meteorologically driven, but also be influenced by biological substances dissolved in the water, such as may tannins, humic acids and chlorophylls, which increase light attenuation through the water column. Gelbstoff (gilvin or ’yellow substance’) is a result of decomposition of organic matter into a complex group of compounds called "humic substances". In general, humic substances are large molecular weight compounds; for example, the average atomic compositionofhumic substancefromtheOkefenokee SwampinGeorgiawas These substances from a molecular of a few hundred to insoluble vary weight macromolecular aggregates. Humic substances absorb light, especially at the blue end of a thespectrum,resultingin shiftinthemaximumpenetrationofspecificwavelengthsoflight Pierce and mineral (Figure 12b; Kirk 1983). et al (1986) found that chlorophyll a,c suspensate accounted for most of the attenuation of light through the water column in the Rhode River Estuary. They suggested that the high attenuation of selected wavelengths in the column upperpartofthewater mayreducetheavailabilityofPARbelowthatnecessary forbenthic plantsorshiftcommunitystructuretofavor speciescapableofusingwavelengths greater than 525 nm. Carter and Rybicki (1990) document a dramatic shift in the quality of light reaching 1 m depth as a result of suspended solids. Natural plankton blooms may also reduce the amount of light transmitted through column. the water Cosper et al (1987) documented blooms of the chrysophyte Aureococcus anophagefferens with cell counts greater than 109 cells L l The bloom of this alga reduced . light penetration through the water column and resulted in the loss of 55% of the eelgrass habitat in Long Island bays (Cosper et al, 1987; Dennison et al, 1989). Texas has recently been experiencing a similar "brown tide" that has been persistent in some portion of Laguna Madre since July 1990. The brown tide organism (BTO) appears to be an undescribed type HI aberrant chrysophyte related toAureococcus anophagefferens and Pelagococcus subviridis. The BTO is a 4-5 pm in diameter; maximum chlorophyll concentrations of 70 Mg L' 1 and densities The BTO has also up to 109 cells L 1 have been reported (Stockwell et al, 1993). had dramatic effects on both the micro and mesozooplankton populations (Buskey and Stockwell 1993). As a result of the high cell densities, light has been dramatically reduced (by up to 60%) in Laguna Madre (Dunton submitted) and could contribute to the loss of habitat. seagrass Banks of drift also influence communities (Cowper 1978; Benz algae may seagrass et al, 1979; Gilbert and Clark 1981; Kulczycki et al, 1981; Zimmerman and Montgomery 1984; Vimstein and Carbonara 1985). The development of the drift algal communities reduce light availability to the seagrasses (Cowper 1978). Additionally, the decomposition of drift algal banks may influence nutrient dynamics (Zimmerman and Montgomery 1984; Vimstein and Carbonara 1985). Although microalgal epiphytes do not decrease water clarity per se, they can influence the amount of light that a macrophyte receives. Numerous studies have documented decreased light availability to macrophytes as a result of epiphyte growth. Some of the earliest work (Sand-Jensen 1977) suggested that diatomaceous epiphytes act as a barrier to carbon uptake and reduce light availability. Bothwell (1989) found that the nutrient to the inner was diffusion limited. supply periphyton layers Additionally, Meulemans (1987) found that light is strongly and selectively absorbed in the upper layers of periphyton communities. Thus, recent studies have confirmed the findings of Sand- Jensen. Other studies have shown that microalgal epiphytes can attenuate 58-94% of light incident on the leaf surface (Batuik et al 1992; Staver 1985; Twilley et al, 1985). ., Anthropogenic influences on water quality and clarity Human activities tend to cause a decrease in overall water quality. For example, enrichment water and estuaries often results in nutrient (decreased quality) in bays that has been used and abused by estuarine eutrophication. Eutrophication is another term and of environmental biologists managers. Eutrophication is a process and community change caused by the interaction of three components: (1) excessive nutrient availability; (2) reduced illumination and (3) a shift in the species composition as a result of altered light and nutrient regimes (Kaldy, 1992). There are numerous investigations from around the world that have shown or suggested that increased nutrient loading stimulate the growth of algal competitors (epiphytes or phytoplankton) which shade out seagrasses (Zieman 1975a; Sand-Jensen 1977; Zieman 1982; Cambridge and McComb 1984; Cambridge et aL, 1986; Zieman and Zieman 1989; Giesen et aL, 1990; Short et aL, 1991; Kaldy 1992; Short et aL, inreview). Thissequenceofeventshasalsobeenshowntoaffectperennialmacroalgaelike and Focus (Kautsky 1991) and freshwater submerged macrophytes like Elodea canadensis Potamogeton pectinatus (Ozimek et aL, 1991). Several conceptual models (Phillips et aL, 1978, Short et aL, 1991, Vogt and Schramm 1991) have been developed to examine the in process of eutrophication (Figure 13; Kaldy 1992, Short et aL, submitted, Kaldy et aL, prep.). Presently, eutrophication problems appear to be confined to a few localized areas of the Galveston and White in Texas, including portions Bay System (Pulich 1991). However, other areas are not immune to eutrophication. Most of the watershed that drains intoCopanoBayissubjectedtoagriculturalfertilizerapplicationthatinfluences thenutrient regimeofthebay(Shormann 1992). Additionally, asignificantamountofthenutrientinput to Copano Bay is from point source sewage outfalls from urban areas within the watershed. Figure 13. Chemical loading hypothesis suggests that under low nutrient levels seagrasses are dominant with some algae present As nutrient loading increases seagrass density decreases and phytoplankton and epiphytes become more prevalent Under conditions of excessive nutrientloadingseagrass densityandbiomassbecomeslow,andoneofthreealgal forms (phytoplankton, microalgal epiphytes or macroalgal epiphytes) becomes dominant. From Short et al 1991. ., Thus, Shormann (1992) suggests that Copano Bay can be considered eutrophic. In addition, recent a considerable amount researchsuggeststhat ofthenitrogeninputtoTexasbaysand estuaries comes from rain falling directly into the bays (Shormann 1992, T. Whitledge pers. comm.). The potential for eutrophication problems along the Texas coast cannot be one of ignored. Reports on population growth rates indicate that the Texas Gulf coast is the most rapidly developing areas of the country (NOAA 1988, 1990a,b). As a result, most of the bays and estuaries along the coast are highly susceptible to pollution, including excessive nutrient loading (NOAA 1990b). Decreased water clarity, due to increased suspended solids, also occurs as a result disturbance Zieman of dredging, construction, erosion, runoff and (Zieman and 1989). Odum (1963) showed that silt from dredging activities in the Gulf Intracoastal Waterway have caused imbalance may an in respiration and photosynthesis resulting in decreased productivity. Zieman (1975a) suggested that dredging was responsible for the destruction of seagrass as a result of direct physical damage, reduced light and hypoxia associated with the high oxygen demand of decomposing material. He also suggested that clam dredges were justasdamagingasdredgingnavigationalchannels. PulichandWhite(1991)suggested that construction, dredging and suspended solids all contributed to the decline of seagrasses in the Galveston Bay system. Giesen et al (1990) suggested that recent large losses of eelgrass from the Dutch Wadden Sea are the result of increased turbidity from both progressive eutrophication and dredging activities. Onuf (1993, in review) found that light attenuation near the dredge spoil site remained elevated 15 months after dredging had occurred. Increased light attenuation near the edge of a seagrass beds was evident for 10 months after dredging. Thus, human construction activities often have a destructive influence on seagrass habitat. also exacerbate water Anthropogenic physical disruption of the environment may Bulthuis et aL (1984) found that the concentrations of suspended solids, quality problems. phosphorus and silicate were higher in water ebbing from denuded mudflats than from covered mudflats. The efflux of nitrogen from the sediments was light mediated seagrass as a result of demand by photosynthetic organisms, and was not different between denuded and covered areas. Bulthuis et al. (1984) suggested that denudation ofseagrass-covered tidal mudflats would lead to increased efflux of suspended solids and nutrients from the sediments to the overlying water. While not generally decreasing water quality, boating activities have deleterious There are in Thalassia effects on seagrasses. several reports of motorboat propeller scars beds (Zieman 1975a, 1976; Dunton pers. comm. 1993). These scars do not recover rapidly after disturbance, and persist for 2-5 years even in healthy, thriving beds (Zieman 1976). In addition to physical damage to the plants the sediment microhabitat is impacted. For example, changes in the grain size, pH and eH of the disturbed sediments have been observed (Zieman 1976). Walker et aL (1989) used aerial and underwater photography to assess the adverse effects of boat moorings on seagrass beds. Moorings scoured circular was 2. patches ranging in size from 3to 300 m Although less than 2% of the seagrass area lost to moorings the increased edge effect makes more of the beds susceptible to erosion and "blowouts". In general, the mechanical disturbance of beds is a worldwide seagrass problem typified by localized impacts associated with construction activity (Short et al.. 1991). Thermalandtoxicpollutionalsohaveadverseeffectsonwaterquality. Ziemanand Wood (1975) have shown that thermal pollution reduces the diversity and abundance of algae and animals near effluent canals. They also suggested that several estuarine plant groupswerelikelytorespondtopollution; seagrasses,macroalgae,phytoplankton,epiphytic microalgae and benthic microalgae (Wood and Zieman 1969). However, seagrasses were relatively more resistant to thermal stress than algae (Zieman 1975a). Tropical seagrasses live close to their thermal tolerance, (e.g. Thalassia testudinum has an optimal temperature rangeof28-30°C);therefore,raisingthetemperatureregimecanbe deleterioustotropical and subtropical estuaries (Zieman 1975a). habitat. Pulich and Toxic pollution has also been implicated in the loss of seagrass White (1991) suggest that pollution, including chemical spills, may have contributed to the decline in the Galveston that of seagrasses Bay System. Livingston (1987) suggested losses in Florida estuaries were mainly the result of decreased water quality from seagrass avarietyofurbanandindustrialsources. Eleuterius(1989)alsosuggestedthatspillsoftoxic substances have contributed to the decline of seagrasses in Mississippi Sound. The consensus researchers is that light is the environmental factor that has among the greatest influence on the depth distribution of Albedo affects the amount seagrasses. of light that actually enters the water and is influenced by latitude and declination of the Earth with respect to the sun. Natural and anthropogenic factors can dramatically alter the amountHumans have oflight reaching seagrasses by increasing scattering and absorption. little, ifcontrol over very any, natural weather phenomena; however, we have the capability to minimize the adverse effects of human activity (Table 5). 48 Table 5. meadows in Florida. list of data concerning historic anthropogenic impacts on seagrass From Livingston 1987. StudyArea Indian River Biscayne Bay Florida Keys Florida Bay Tampa Bay system Charlotte Harbor Pensacola Bay system Choctawhatchee Bay St.Andrews Bay St. Joseph Bay Apalachicola Bay system 4 Apalachee Bay Location Southeast Florida Atlantic Ocean Southeast Florida Atlantic Ocean South Florida Atlantic Ocean South Florida Southwest Florida Gulf of Mexico Southwest Florida Gulf of Mexico Northwest Florida Gulf of Mexico Northwest Florida Gulf of Mexico Northwest Florida Gulf of Mexico Northwest Florida Gulf of Mexico North Florida Gulf of Mexico North Florida Gulfof Mexico Status of SeaMeadows grass Historic declines in number and of coverage meadows. Declines in Veto Beach seagrass area, Fort Pierce Inlet (25%) and Sebastian Inlet (38%) from 1951 through 1984. Undetermined deterioration in northern Biscayne Bay. Some damagetoThalassia-Halodule beds near power plant (heatedeffluents) in south Biscayne Bay. Card Sound unaffected by power plant discharge. Few data found. Little effect of Key West desalination plant. Postulated altered species relationships due to increased salinity caused by redirection of freshwater runoff. Almost forty percent reduction in Boca Cicga Bay due to dredging, filling, and associated activity from 1950 through 1968. Multiple sources (urbanization, storm water runoff, sewage discharge, industrialization,toxic substances). Reduction of seagrass meadows in Tampa Bay system from 30,970 ha to5,750 ha. Decline of 29 percent of seagrass beds from 1943 through 1984. Complete loss of seagrass beds in Escambia Bay, East Bay, and Pensacola Bay from 1949-1979. Some fresh-brackish water species extant in delta areas. Some Thalassia-Halodule beds still alive in Santa Rosa Sound. Losses due to urbanization, industrial waste discharge, dredgingand filling, cultural eutrophication. Historical deterioration of seagrass beds from 1949 through 1983. Causes unknown. to No data found: Presumed impact cue urbanization, industrialization Extensive coverage unchangedfrom 1972 through 1983. Relatively unpopulatedarea. Generally healthyassemblages ofseagrasses. Localimpact duetodredgedopeningin associated barrier island. Introduced species spreading in delta areaswith as yet undetermined impact. Area under increased pressure from urganization. Impactsduetodisposalofpulpmill wastes (Fcnhollowayestuary)from 1954tothepresent. Slow recovery noted inouterportionsofimpact area (associated with pollution abatement program). Area now threatened by proposed inshore navigation channel and possible off­shore oil drillingoperations. Information Source Goodwin and Goodwin, 1976; Florida Department of Natural Resources, unpub­lished data. McNulty, 1961; Roessler and Zieman, 1969; Thorhaug et al., 1973; Zieman, 1970, 1982. Chcshcr, 1971 Zieman, 1982 Lewis and Phillips, 1980; Simon, 1974; Lewis et al., 1985; Taylor and Saloman, 1968 Harris etal., 1983 Livingston, 1979; Livingston etal., 1972;Olingcretal., 1975 Burch, 1983 McNulty et al., 1972; Savastano et al., 1984. Livingston, 1980c, 1983 Heck, 1976; Hooks et al., 1976; Livingston, 1975, 1982a, 1984a; Zimmerman and Livingston, 1976a,b. 49 CHAPTER 3: THE CHESAPEAKE BAY APPROACH Approaches to assessing habitat requirements of seagrasses Recently, several documents have been published which present a novel approach to aL, 1992; Dennison et aL, 1993). Dennison the management ofseagrass habitat (Batuik et et aL (1993) presents the major findings of a comprehensive technical synthesis conducted by Batuik et aL (1992). Both Batuik et al (1992) and Dennison et al. (1993) advocate the measurement of physical and chemical water column parameters in existing seagrass beds as a means of identifying impacted systems. The premise is that by knowing the conditions required for plant growth, we can determine the water quality of a specific area through examination of or absence. Thus, areas indicate seagrass presence where plants do not grow that some water quality parameter(s) do not meet the minimum requirements of the plant. Waterqualityparameterrequirements wouldbespecificforeachplant speciesandestuarine habitat. Chesapeake Bay parameter requirements were developed by monitoring water quality gradients within the system over time (Batuik et aL, 1992). The habitat requirements developed in this manner represent the absolute minimal water quality characteristics necessary to sustain plants in shallow water (Dennison 1993). The growth, survival and depth distribution ofsubmerged aquatic vegetation (SAY), including seagrasses, is related to underwater light availabilty (Chapter 4). The parameters examined were a to determine habitat requirements total suspended solids, chlorophyll levels, dissolved inorganic nitrogen and phosphorus and the light attenuation coefficient (Batuik et aL, 1992; Dennison et aL, 1993). These parameters affect light availability in a variety ofways, either by directly absorbing and scattering light or by stimulating the growth ofphytoplankton(seeChapter2forfurtherdiscussion). Thehabitatrequirementapproach does on not rely on understanding the interactions of water quality and light but relies empirical water quality data and SAY survival (Dennison et aL, 1993). Measurements were of the water quality parameters made monthly through the growing season, although longer databases (up to 10 years) do exist for some portions of the Chesapeake Bay System. Minimum habitat requirements for SAY in polyhaline areas (salinitygreaterthan 18%o)wasalightattenuationcoefficientof1.5(m'1),totalsuspended solids of 15 mg L l, chlorophyll aof 15 /xg L 1 and dissolved inorganic nitrogen and phosphorousof10and0.67/xM,respectively. Otherminimumrequirementsweredeveloped for other salinity regimes (Table 6). In areas where the water quality parameters do not exceed these values one would expect to find SAY. Although the habitat requirement approach is unique and potentially very useful, there are some drawbacks. Measurements were carried out monthly during the growing season. Due to the immense spatial and temporal variability of marine systems, monthly sampling of water quality parameters is not adequate to characterize a system. Monthly sampling is also likely to underestimate parameter levels experienced by SAY due to the fact that most field avoid sampling during inclement weather. while programs Also, sampling during the growing season may be appropriate for plant species that overwinter as a tuber or seed, it may not be appropriate for species that grow throughout the year (i.e. Zostera marina). In addition, estimates of the light attenuation coefficient are derived from Secchi depths (Batuik et aL, 1992; Dennison et aL, 1993). In view of the problems related Table 6. For each Chesapeake Bay submersed aquatic vegetation habitat requirements. parameter, the maximal growing season median value that correlated with plant survival is given for each salinity regime. Growing season defined as April-October, except for tidal fresh = polyhaline (March-November). Salinity regime defined as 0.05%0, oligohaline = 0.0-5%0, mesohaline = 5-18%0, polyhaline more than 18%0. From Dennison et al 1993. ., Light Total Dissolved Dissolved attenuation suspended inorganic inorganic coefficient solids Chlorophyll nitrogen phosphorus Salinity regime (mg/l) a (Mg/l) (mM) (mM) Tidal freshwater 2.0 15 15 0.67 _ — 2.0 15 15 0.67 Oligohaline Mesohaline 1.5 15 15 10 0.33 Polyhaline 1.5 15 15 10 0.67 to the use of Secchi depth to measure attenuation (see Chapter 2) it seems that direct measurement of light availability is more appropriate. The habitat requirement approach may be useful tool to estuarine managers in a It developing water quality standards to prevent the loss of SAV, including seagrasses. should be pointed out that there is a fundamental difference between the Chesapeake Bay and the bays and estuaries in Texas. The Chesapeake Bay program is using water quality criteria for the reestablishment and restoration of a highly perturbed system. In Texas, the goal is to establish water quality criteria to prevent the destruction of habitats. seagrass Consequently, a different approach and strategy may be appropriate for preserving seagrass habitats in Texas. CHAPTER 4; SEAGRASS PHOTOSYNTHETIC PHYSIOLOGY Historically, muchoftheworkonthephotosyntheticphysiologyofmarineplantshas been done with phytoplankton, as they are at the base of the oceanic food web. Based on these studies, the euphotic zone was defined the depth to which 1% of surface PAR as penetrates (Bougis 1976; Raymont 1980). The 1% light level is not appropriate for defining the depth limits of seagrasses due to the higher respiratory demands of the below ground tissues (Kenworthy and Haunert 1991). Until recently, it was generally accepted that 10% of surface PAR was required to sustain seagrass populations. More recent studies suggest that most seagrass species require 15-25% of surface irradiance (Table 7; Kenworthy and Haunert1991;DennisonetaL, 1993).Numericalmodelsof seagrass depth limits have been developed and used to predict the depth distribution of submerged macrophytes. Many of Chambers the models are regression models of field data using least squares methodology. andKalff(1985)developedregression modelsforavarietyoffreshwatermacrophytesfrom Canada (Table 8). They used attenuation coefficients developed from mean summer Secchi depth to estimate light availability. In general, approximately 20% of surface PAR was required to sustain freshwater macrophyte species. Duarte (1991) developed a regression model (Table 8) that predicts depth distribution for a variety ofseagrass species. According to his model, most seagrasses require about 11% of surface PAR (Duarte 1991). Many of the species used in Duarte’s (1991) model occur in clear waters and very thus may underestimate the minimum light requirement for In addition to this seagrasses. problem,bothChambers andKalff(1985) andDuarte (1991)relyonattenuationcoefficients Table 7. were calculated Maximaldepthlimitsandminimallightrequirementsofvariousseagrassspecies. Minimallightrequirementsaspercentlightatmaximaldepth. Ranteofmaximaldepthlimit andmeantSEofminimallightrequirementsgivenforlocationswith multiple data points. From Dennison etal., 1993. Maximal Minimal light Genus and species Location depth limit (m) requirement (%) 7.0 Amphibolisantarctica* WaterlooBay(Australia 24.7 4.0 10.2 Cymodocea nodosa Ebro Delta (Spain) Malta 38.5 + Halodule wrighdi Florida (US) 1.9 17.2 Halophila decipiens^ St. Croix (US) 4.4 C. nodosa 7.3 40.0 Northwest Cuba 24.3 8.8 Northwest Duba H. decipiens 14.4 23.7 Heterozostera tasmanica 3.8-9.S 5.0±0.6 Halophila engelmanni Victoria (Australia) Chile 7.0 H. tasmanica 17.4 H. tasmanica 39.0 4.4 Spencer Gulf (Australia) H. tasmanica 8.0 20.2 Waterloo Bay (Australia) Poisdonia angustifolia Waterloo Bay (Australia) 7.0 24.7 Island (Spain) Poisdonia oceanica Medas 15.0 7.8 P. oceanica Malta 35.0 9.2 Poisdonia ostenfeldii Waterloo Bay (Australia) 7.0 24.7 Poisdonia sinuosa 7.0 24.7 Waterloo Bay (Australia) Ruppia maritima Brazil 0.7 8.2 Syringodium filiforme Northwest Cuba 16.5 19.2 S. filiforme* Florida (US) 18.3 6.8 + Florida (US) 1.9 17.2 Thalassia testudinum Northwest Cuba 14.5 S. filiforme 233 T. testudinum Puerto Rico 1.0-5.0 24.4±4.2 T. testudinum Florida (US) 73 15.3 Zostera marina Kattegat (Denmark) 20.1±2.1 3.7-10.1 Z marina 2.0-5.0 19.4±1.3 * Roskildc (Denmark) Zmarina Denmark 13-9.0 20.6H3.0 Z marina Woods Hole (US) 6.0 18.6 Z marina Netherlands 23 29.4 Z marina 2.0-5.0 Japan 18.2±43 *Duartc 1991 + WJ. Kenworthy, personal communication, 1990 and Dennison 1990 S Ostenfcld 1908 #Borum 1983 55 Table 8. Regression equations used to predict the depth distribution of different aquatic plants. Group Equation Freshwater = Angiosperms (ZJ-3 1.33 log(D) + 1.40 Bryophytes (Zys = -0.48 log(D) + 0.81 c = Charophytes log(Z) 0.87 log(D) + 0.31 c Marine Seagrasses = - general log (Z) 0.26 1.07 log(k) c - Thalassia Z = 0.27 0.93 log(k) c - Zostera Z = 0.27 0.84 log(k) c Chambers and Kalff 1985 2From Duarte 1991 Z = colonization depth c D = Secchi depth mean summer k = diffuse light attenuation coefficient from Secchi measurements. Priesendorfer that it is developed depth (1986) suggests notappropriate to derive attenuation coefficients from Secchi depth (see Chapter 2). Computer modelling Computer simulation models of seagrass growth and productivity in relation to light and nutrients have been developed. The development of mathematical models provide a mechanism for synthesizing the information available in the literature (Short 1980). Models have become predictive tools for management, but can also provide insight to the dynamic of One of the first mathematical models of response seagrass ecosystems (Short 1980). how nutrients affected seagrass ecosystems investigated eelgrass growth; however, to realistically model the system it was necessary to incorporate light limitation (Short 1980). ShortutilizedSteele’sequationforphytoplanktonphotosynthesis, whichdescribesproduction as increasing with increased light up to an optimum light intensity. Beyond the optimum, as a production decreases result of photoinhibition (Short 1980). The model was run using data from Charlestown Pond, R.I. The simulation yielded a good representation of the seasonal trends and a reasonable fitto the observed data (Short 1980). Wetzel and Neckles a (1986) developed model of photosynthesis and growth for Zostera marina in relation to selected physical-chemical variables. They found that physical parameters such as light and temperature controlled growth and photosynthesis. Small changes in submarine irradiance or temperature resulted in decreased plant productivity and the eventual loss of the seagrass community. These simulations suggest small changes in light (or temperature) may cause the complete loss ofseagrass communities on the edge of their physiological tolerance (e.g., Wetzel and Heckles (1986) concluded that ambient the deep edge of the seagrass bed). light was a principle factor controlling the longevity and survival of seagrass beds. Zimmerman et al (1987) modelled nitrogen budgets and light availability using field and were tested experimental data from numerous investigations, and the model predictions against other field data. They found that light had a significant effect on the rate and site marina. The to H of nitrogen uptake in Z. model predicts that for eelgrass exposed sat greater than 6 h (i.e., "normal" conditions) most of the nitrogen uptake will occur through the roots. However, in low light environments (short there is an increase in the importance of nitrogen uptake and assimilation by the leaves. Thus, the site of nutrient uptake appears to be partially dependent on the light environment the plant experiences. As a general approach to seagrass ecology, simulation models are invaluable. They can be used to synthesize the available data and point out new directions for research. that mathematical simulations could be used Short (1980) showed to adequately predict seagrass productivity as a function of several parameters including nutrients, light and Wetzel and Heckles that small in the diffuse temperature. (1986) suggested changes attenuation coefficient could result in the elimination of habitat. Zimmerman et seagrass the al (1987) suggest that the site of nutrient acquisition is dependent, at least in part, on light environment the plant experiences. In general, simulation models can be very interesting and provide useful information to seagrass biologists and estuarine managers. Effects of in situ light reduction Although computer simulations are informative, researchers also need to quantitatively evaluate the effect of model parameters on in situ populations. During the late 1960’s and studies were conducted to examine the effects of reduced 1970’s numerous light on seagrass growth, production and morphology. Some of the earliest work involving was done by Burkholder and Doheny (1968), who used to reduce seagrass shading cages available 10 and 1.6 % of surface irradiance (SI). Eelgrass (Zostera light to 100, 60, 20, marina) growing in these cages became noticeably stunted and did not survive at light levels less than 20% SI. In an attempt to validate a model, Short et al (1974) investigated how the hydrodynamics of the Charlestown Pond R.I. were influenced by eelgrass. They examined the influence of light on seagrasses (and in turn the hydrodynamic model) by which resulted in shorter plants as well as reduced biomass and shading eelgrass with cages, density. Backman and Barlotti (1976) reduced 63% of the light reaching Z. marina in a coastal lagoon in California for a period of nine months. They found reduced biomass, shoot density and incidence of flowering. Congdon and McComb (1979) examined the reduced light in Australian estuary. They an productivityofRuppiamaritimainresponse to used seven light levels in the study: 100, 60, 41, 28, 19, 15, and 7.5% SI. However, they made no in situ measurements of irradiance. They reported a general seasonal pattern of As the duration of low standing crop during winter with a rapid increase in the spring. shading increased, the plants required higher light levels to persist. At least 20% SI was for requiredtomaintain50%ofinitialstandingcrop upto100days,whilegreaterthan60% SIwasrequiredtomaintain50%ofinitialbiomassformorethan200days. Theyconcluded thatareductioninlightintensitymayresultinthelossofconsiderable quantitiesofRuppia. Duringthe 1980’sseveralotherinvestigationsexaminedtheinfluenceofreducedlight various Bulthuis (1983a) examined the of on seagrass species. For example, response Australia. Heterozostera tasmanica to in situ light reduction in Victoria, The light levels below the screens were approximately 18, 13, 4.7 and 1% SI. Light levels less than 4.7% SI resulted in the death of all shoots within 2to 10 months. Light levels of 18 and 13% SI caused reduced shoot density relative to the controls, suggesting the plants were unable to survive indefinitely at these light levels. Some changes in shoot morphology were also noted;however,leafgrowthrateandleafwidthremainedthesame. Bulthuis(1983a)data indicate that H. tasmanica have may a higher light requirement at summer temperatures than at winter temperatures and thus may be more sensitive to reduced irradiance during summerthan in winter (Bulthuis 1983a). Carter and Rybicki (1985), found that both light penetration and grazing pressure effected the survival of transplanted Vallisneria americana Odum (1985) examined the of Thalassia in the tidal Potomac River, Maryland. response testudinum to shading by epiphytes. She did an in situ shading experiment during the winter where light intensity was reduced to 2-10% SI. Shoot density decreased under the reduced light treatments and after eight months all the plants were dead. It was suggested that demise of the would have occurred more during summer the seagrasses quickly. Neverauskus (1988) examined the response of Posidonia sinuosa and P. angustifolia in Australiatochroniclong-termlightreduction(Figure 14). Heconstructed acanopyof50% shade cloth, which was placed over the community. During the first six months of treatment, shoot density was unchanged but leaf density and standing crop had declined; during the second six months shoot density decreased dramatically. However, although Neverauskus (1988) documented changes in plant canopy structure, the study itself was very poorly Figure14. Changesinstandingcrop,leafdensityanddryweightofepiphytesofPosidonia in response to 50% reduction in ambient light. Error bars represent standard deviation, n = 4. From Neverauskas 1988. designed. There was no control plot, no statistical analysis nor was there any estimate of the actual amount of light received by the plants. Stable carbon have been to examine the effects of isotope ratios used recently 13 reduced light on The 1991). Carbohydrate reserves may act as a buffer during periods when H is less than that required by the ml plants; however, the reserves are limited in nature. Plants at the lower edge of the beds may not be able to build up adequate reserves to sustain them through periods of low light. Thus, brief periods of extreme turbidity may be more critical than the mean turbidity in controlling the depth distribution of existing populations and the establishment ofseedlings or as well as the propagules. The average length of the daily H period is important, sat number of "critical days" per month or season when H requirements are not met. The sat number of extreme attenuation days (BAD) when diffuse attenuation values prevent net of habitat carbon gain or adequate root oxygenation may provide a quantitative measure tobe in suitability that may prove a sensitive predictor of eelgrass growth and survival different habitats (Zimmerman et aL, 1991). Laboratoiy studies of reduced light on seagrasses In addition to in situ studies of reduced irradiance and photoperiod there have been two in vivo studies the to reduced examining response of submerged aquatics light. and examined the Goldsborough Kemp (1988) response of Potamogeton perfoliatus to changes in total irradiance. Plant populations were grown in aquaria and the light environment was manipulated using neutral density screens which alter light intensity but The 100% SI. not spectral quality. treatment levels were 11, 32 and After three days, shaded plants showed an increase in photosynthetic efficiency and chlorophyll a concentrations. Following ten days of treatment there were significant changes in the morphological characteristics of the plants including elongation of stems, thinning of lower leaves, and canopy formation at the water surface. Golsborough and Kemp (1988) concluded and conferred that the physiological morphological responses of the plants improvements in plant fitness under the treatment conditions. Tomasko (1992) examined changesinthemorphologyofHalodulewrightii duetochangesinthespectralcomposition of light. Halodule was grown under a canopy of Thalassia testudinum that changed the ratio of red;far red light. Other plants were grown under neutral density screens of equivalent light reduction that did not alter the redifar red ratio. Plants under Thalassia grown testudinum had longer internode lengths, while plants under neutral density screens showed reduced growth rates compared to controls. Tomasko (1992) suggested that Halodule minimizes competitive interaction with Thalassia by varying its morphology. Studiesofreduced irradiance and dailylight periodsuggest thatwhen seagrasses are light stressed they change their morphology (within limits) to optimize photon capture. Thus, it appears that morphological plasticity may confer some advantage to submerged reduced irradiance. macrophytes, allowing them to survive periods of The work examining of eelgrass and the laboratory studies of Potamogeton also suggest these plants may physiologically adapt to reduced light. curves Photosynthesis irradiance Photosynthesis irradiance (P vs. I) curves are a means of quantifying the Blade tissue is photosynthetic rate on an area, weight or chlorophyll basis (Kirk 1983). placed in a chamber with an oxygen electrode and a light sensor. Oxygen evolution or a In consumption is measured as function of light intensity striking the plant (Figure 16). as a result of the dark there is no photosynthesis and plants exhibit net consumption of0 2 respiration. As light intensity is increased, some O production occurs; however, this shows z up as a diminution of the 0 2 consumed by the blade tissue (respiration exceeds irradiance 0 the 0 photosynthesis). The at which photosynthetic 2 production equals 2 consumed in respiration is the compensation irradiance (/). At light intensities greater than c I the plants exhibit net photosynthetic production. Maximum photosynthetic production c (P) is achieved when increases in PAR no longer result in an increase in oxygen max be estimated from the intersection of the initial evolution;thelightsaturationpoint(Ik) can slope with Pmax or is more accurately calculated as P /0.. Alpha (a) is represented by the max initial slope of the P vs. I curve and is a measure of the efficiency with which the plant biomass utilized light (Kirk 1983). It represents the efficiency with which light quanta are 11. reflects the absorbed and used to transfer electrons through photosystems I and dark reactions of the various metabolic of photosynthesis and is regulated by recycling intermediates (ATP and NADPH). Although there have been numerous studies of the P vs. I parameters of seagrasses, few have been carried out in situ. Wetzel and Penhale, (1983) examined in situ the P vs. I characters that allowpopulations ofSAY inChesapeake Bay tosurvive inavery stochastic Figure 16. light saturation curve for photosynthesis. P, photosynthesis; P maximum max, incident photosynthesis; gross photosynthesis; P net photosynthesis; R, respiration; n, I PAR. From South and Whittick 1987. PAR; saturating PAR; compensation k, environment. They found that submarine PAR and temperature act both singularly and control the biomass and distribution of For interactively to seagrasses in Chesapeake Bay. maritima P correlated with and of Ruppia was temperature there was no sign , max photoinhibition. P for Zostera marina was not correlated with temperature, although max Zostera appears to have a temperature optima below 28 °C. Wetzel and Penhale (1983) found that the two species were therefore physiologically distinct, allowing them to coexist in the same niche. The authors suggested that changes in the distribution of in seagrasses Chesapeake Bay may be related to changes in light availability. Libes (1986) examined the P of Posidonia oceanica in situ vs. I relationship using Cl4 techniques. Photosynthetic summer. efficiencywasgreatestinwinterandleastin Additionally,productivitywasgreatest in the morning and least in the evening. Libes also found evidence of seasonal photoinhibition and suggestedthatPosidonia mayhaveaseasonalendogenousrhythmwith regard to photosynthetic capacity. Ic andIk were similar for both the seagrass and its epiphytes. Dunton and Tomasko (submitted) made in situ measurements of the photosynthetic performance of Halodule wrightii in Texas. They documented seasonal variation in all of the parameters that describe the P vs. I curve. Laboratory studies of P vs. I characters are easier to perform than field studies, and as a result, numerous laboratory studies of seagrass P vs. I characteristics have been completed. Drew (1978) investigated the factors affecting photosynthesis and its seasonal variation in Cymodocea nodosa and Posidonia oceanica. He measured photosynthesis, dark respiration and leaf chlorophyll content on plants from shallow (1-5 m) and deep (25-33 m) sites. Cymodocea had similar spring and summer light saturated photosynthetic rates, while Posidonia had higher photosynthetic rates in spring than in summer (Figure 17). The for photosynthesis was about 30 °C for both species and dark temperature optima respirationwas similar during both the spring and summer. Drew (1979) examined the physiological aspects of primary production in Cymodocea nodosa, Posidonia oceanica, PI Halophila stipulaceaPhyllospadix torreyi, Zostera angustifolia and Zostera marina. vs. , curves were developed for each species (Figure 18) as were plots of photosynthesis versus temperature. All species showed similar P vs. I curves that usually shows light saturation around 2-3 mW/cm2 (~138 pmol m'2 s'1). Halophila was the only species that showed evidenceofphotoinhibition. DespiteawiderangeofP (8.1to26.0/igCcm'2h'1)allof max the Ik values were about 10% of full sunlight and all of the Ic values were about 1% SI (Table 11). Williams and Mcßoy (1982) examined the effects of light on carbon uptake in six seagrass species: Thalassia testudinum , Syringodium filiforme, Halodule wrightii, Halophila engelmanniPhyllospadix scouleri and Ruppia maritima. Seagrasses from Texas became light , saturated between The five had similar half- at high irradiances, 64-85% SI. species saturation constants ranging from 49-56% SI. Williams and Mcßoy (1982) suggested that Thalassia and Syringodium are "climax species" and that the other species are "colonizers". to Kerr and Strother (1985) examined how photosynthesis in Zostera muelleri responded irradiance, temperature and salinity (Figure 19). Photosynthesis increased with increasing light (from 17 to 185 fimol m'2 s* 1) at 16 °C, with no indication of photoinhibition at the highest light levels used in the experiment. Because thedarkreactionsofthephotosynthetic process areenzymaticallymediated. Figure 17. Effect of temperature on net photosynthesis and dark respiration in shallow-growing leaves ofCymodocea and Posidonia in spring and summer; photosynthesis measured 2 at 20 mW cm' From Drew 1978. . Figure 18. P vs. I curves for six seagrass species at ambient environmental temperatures From Drew 1979. Table 11. Photosynthetic rates, respiration rates, saturation and compensation irradiance at ambient environmental temperatures in six seagrass species. From Drew 1979. RR Ymax \ Ic Temp. (*C) ic Phyllospadix torreyi 26.0 3.6 0.5 -3.8 -5.3 15 21.8 3.8 0.4 -2.5 -3.5 25 Zostera angustifolia 18.i 3.2 0.3 -1.8 -2.1 10 Zostera marina 14.2 5.0 0.6 -1.7 -1.4 15 Cymodocea nodosa 9.0 2.0 0.2 -1.2 -1.1 25 Halophila stipulacea Posidonia oceanica 8.1 2.6 0.4 -1.3 -1.5 17 22 2 *yMgCcm'h1;Iand7mWcm'PAR;R}-andR /xgCcm' ]h-1. max* tt C, d, Figure 19a. Relationshipbetweenapparentphotosynthesisandirradiance(PAR)inZostera muelleri at 16 °C. From Kerr and Strother 1985. 19b. Figure Relationship between apparent photosynthesis and temperature in Zostera muelleri at an irradiance of 47 ± 2 /xmol m‘2 s'1 From Kerr and Strother 1985. . effect the Several temperature can have a pronounced on (Bulthuis 1987). ofP vs.Icharacters investigations have examined the response to changes in temperature. examined As part of his study, Drew (1979) the effects of temperature on the rates of photosynthesisandrespiration. Measurementsofphotosynthesisandrespirationweremade at six temperatures between 10 and 40 °C (Figure 20). Photosynthesis increased linearly with increased temperature {r2= 0.99 to 1.00). For some species (Posidonia and Cymodocea) temperatures greater than 30 °C caused thermal damage and resulted in decreasedphotosynthesis. Attemperatureslowerthannormallyencounteredintheirnatural a habitat, all species maintained moderately low rate of respiration. Bulthuis (1983b) on the P vs. I curve of Heterozostera tasmamca. He examined the effects of temperature At measured the P vs. I parameters at 8 temperatures ranging from 5 to 40° C (Figure 21). ° 25 and 30 C photosynthesis was not light saturated even at the highest intensities (955 /xmol m*2 s'1). P increased by a factor of 2.5 between 5 and 30 °C and decreased sharply nmx between 30 and 35 °C. factor of 2 between 5 The rate of dark respiration increased by a and 20 °C with small increases above 20 °C. Evans examined (1984) the physiological response of Chesapeake Bay populations of Zostera marina and Ruppia maritima to temperature. For both species the lowest P was at 8° C. The highest occurred for max °° Zostera at 19 C and for Ruppia at 26 and 30 °C. At temperatures between 8 and 19 C 2° Zostera had a higher P than Ruppia. I for Zostera ranged from 46 pmol m' s'1 at 26 C max k 21 to s’ 28pmolm’at30°C,whileforRuppiatherangeatthesametemperatureswasgreater (39-72pmol m2s'1). Ingeneral,RuppiahadahigherI thanZosteraatalltemperatures. k Evans (1984) suggested that at lower temperatures Zostera has the competitive edge over Figure 20. Effect of temperatures both above and below ambient on the rates of gross in four Horizontal bar is the of environmental photosynthesis seagrasses. range temperatures normally encountered by the plant; open circles reflect the incubation. From Drew 1979. Figure 21. Temperature vs. apparent photosynthesis in Heterozostera tasmanica at three irradiance levels. Mean ± 1 SE (n = 3-9) for plants collected from Crib Point, Australia. From Bulthuis 1987. Ruppiawhile at higher temperatures Ruppia has the advantage. Thus temporal resource , partitioning may allow the coexistence of these species. Kerr and Strother (1985) examined the photosynthetic rate of Z. muelleri at light levels of 47 fimol m 2 s'1 and a range to 30 °C and showed a oftemperatures. Photosynthesis increased with temperature from 5 marked decrease at temperatures greater than 30 °C (Figure 19). Marsh et aL (1986) examined the effects of temperature on photosynthesis and at 20-22 °C respiration in Zostera marina. They developed P vs. I curves for plants grown and subjected to various temperatures between 0° C and 35 ° C (Figure 22). Light saturated net photosynthesis increased with temperature to an optimum around 25-30 °C and decreased at 35 °C. The initial slopes of the P vs. I curve (a) were greatest at 0° C and least at 35 °C. Kerr and Strother (1985) suggest that in winter, low temperatures and low light conditions enable the plants to maintain a positive carbon balance; however, high Bulthuis temperatures (>3O°C) and low light would result in a negative carbon balance. (1987) reviewed the effects of temperature on seagrasses. He found compensation irradiance increases with temperature; therefore, to maintain a positive carbon balance, seagrass plants require a greater irradiance during summer than during winter. The literature indicates temperature optima for light saturation of photosynthesis is generally between 25 and 35 C; however, there tends to ° be a rapid loss of photosynthetic capacity at I temperatures above the optimum. Bulthuis (1987) suggests that the seasonality of may c at near have important implications for seagrasses living or their minimum light requirements. Madsen and Adams (1989) found that Potamogeton pectinatus exhibited ° optimalphotosyntheticproductionat30 °Candthatphotosynthesisat10 Cwas63%lower Figure 22. P vs. I relationships of Zostera marina leaf segments measured at different temperatures. Leaf tissue was grown at 20-22 °C in the field was incubated at the given temperatures for 15 minutes prior to measurement; a, photosynthesis expressed per mg Chlor.'1min'1;b,photosynthesisexpressedperdm*2min'1 FromMarshetal.,1986. . of than that at 30 °C. Perez and Romero (1992) examined the photosynthetic response Cymodoceanodosatolightandtemperature. ThePvs.Icurvesshowedseasonalvariation, with and Ik higher during summer than during winter(Figure 23). I did not exhibit any c seasonality. Cymodocea showed no indication of photoinhibition at light levels up to 2500 nmo\ m'2 s'1 Rates of light saturated photosynthesis and dark respiration increased . significantly with increasing temperature. Perez and Romero (1992) also suggested the control of be seagrass seasonality may more complex than previously thought. As with most other fields of science there are some controversies with regard to seagrassesandphotosynthesis. Oneofthebiggestdifficultiesisstandardizingunitsoflight energy. Through the years a bewildering variety of units have been used to measure light energy, everything from Langleys (Williams and Mcßoy 1982) to milliWatts cm'2 (Drew 1979) to /xEinsteins (Kerr and Strother 1985, Perez and Romero 1992). Although there are numerous conversion factors in the literature, most do not take into account the selective absorption of specific wavelengths of light by water (Liming 1990; Megard and Berman, 1989). Thus, most calculations that employ conversion factors between units are relatively poor estimates of light energy at depth. The best method for estimating underwater light availability is to measure itin situ (Dunton, submitted). One of the other controversies in the literature is the validity and/or usefulness of carbon budgets based on P vs. I curves developed from leaf segments (Fourqurean and Zieman 1991a). TheI valuedevelopedforleaftissuedoesnottakeintoaccountthelarge c respiratory demands of the below-ground tissue. Thus, whole plant carbon budgets based on leaf tissue probably underestimate the true carbon budget of the whole plant Figure 23a. Seasonal trends in P vs. I curves in Cymodocea nodosa for November (boxes), From Perez and Romero 1992. June (triangles), and August (circles). Figure 23b. Response of both net photosynthesis (open symbols) and dark respiration (black symbols) to temperature in August (triangles) and February (circles). From Perez and Romero 1992. 84 (Fourqurean and Zieman 1991 b). Fourqurean and Zieman (1991a) used a P vs. I chamber that allowed them to develop P vs. I curves for whole plants in a more natural orientation to the light field. They found that leaves account for less than 50% of the total plant respiration. Knowledge of the above-to below-ground apportionment and the factors that control this ratio are critical to modelling carbon budgets of seagrasses. In addition, the leaf tissue P I data in models of plant carbon vs versus question of using whole plant budgets must still be addressed. Measurementsofphotosynthesisinseagrasses suggestthatlightandtemperatureare thefactorsthatcontroltherateofphotosynthesisandrespiration. Ingeneral,photosynthesis at a given light level increases until the optimum temperature is reached. At temperatures above the optimum, photosynthesis decreases. Compensation irradiance also varies with as a increased temperature result of increased respiration. Thus, at elevated temperatures plants require greater irradiance to maintain a net carbon balance. The observed seasonal result of seasonal pattern in P vs. I parameters is probably a temperature changes. Seasonal changes in seagrasses Seagrasses, like most plants, exhibit seasonal changes in production, P vs. I organic constituents. Odum (1963) examined parameters and the productivity of Texas Thalassiatestudinum andreportedamarkedseasonalcycleinproductivitywithconsiderable interannual variation (Figure 24). Zieman (1975b) examined the seasonal variation of Florida Thalassia with reference and Maximum to temperature salinity. productivity, standing crop, leaf length and blade density occurred by early summer. He found that 1957-1961 for Figure 24. Record of salinity, gross photosynthesis and total respiration Thalassia testidinum in Texas. From Odum 1963. although Thalassia undergoes marked seasonal variation, blade density during winter was about 50% of maximum blade density during summer. Maximum blade density occurred during late April and May and leaves were carried until October. Minimum blade densities occurred from December through March. Zieman (1975b) concluded that temperature and in Florida. salinity were the major factors controlling the seasonal variation of Thalassia Walker and McComb (1988) examined the seasonal variation oiAmphibolis antarctica and Posidonia australis in Shark Bay, Western Australia. Posidonia showed no clear seasonal or stock for the duration of the pattern of production standing study (1982-1983). coincident with highest light and Amphibolis had maximum production during summer, temperatures. Although seasonal variation was evident for Amphibolis, it was not very pronounced,whilePosidonia didnotexhibitaseasonalpattern. Ingeneral,itappearsthat seagrass biomass and density increase with increasing light and temperature during late spring and early summer (Figure 25). Dunton (1990) examined the production ecology of Ruppia maritima and Halodule wrightii in two estuaries along the Texas coast. The species time and of differed with respect to seasonality of growth, of flowering persistence was a overwintering populations. In both estuaries Ruppia strict opportunist colonizing bare no areas yearly (i.e. overwintering populations) and completing its growth cycle in four months. Halodule was absent from the Guadalupe Estuary (San Antonio Bay), but in the Nueces Estuary (Corpus Christi Bay) itproduced overwintering populations with year-round growth. Dunton (1990) concluded that in some Texas estuaries Ruppia grows as an annual weed, whereas Halodule grows as a perennial (Dunton 1990). However, other investigators have documented perennial populations of Ruppia maritima in Upper Laguna Madre and Figure 25a. Some seasonal environmental features of an Australian study site: total irradiance (squares, Wm’2), PAR photon flux density (triangles, fimol m'2 s’1) and ° temperature (C, circles). Values are monthly averages. From Perez and Romero 1992. Figure 25b. Seasonal pattern of above-ground biomass. Standard errors are indicated by vertical bars. From Perez and Romero 1992. 88 Redfish Bay (Pulich 1985). Thus, both annual and perennial stocks ofRuppia occur along the Texas coast. have noted that the P vs. exhibit a Many investigators I parameters of seagrasses seasonal examined factors affecting the seasonal variation in pattern. Drew (1978) photosynthesis of Posidonia and Cymodocea in the Mediterranean. Cymodocea had while Posidonia had higher comparable photosynthetic rates during both spring and summer, summer(Figure17). Calculationofacarbon photosynthetic rates during spring than during balance based on the experimental P vs. I numbers shows that shallow Posidonia a community may have a positive carbon balance during spring and a negative carbon balance (due to higher respiratory demands) during summer (Drew 1978). However, it should be noted that the carbon budget calculations based on leaf P vs. I curves do not take into accounttherespiratorydemandsofthebelow-groundtissues. CongdonandMcComb(1979) found thatRuppia exhibits seasonal changes in above and below-ground biomass, with low crops occurring during late autumn and winter. Although they did not investigate seasonal changes in the P vs. I curves per se they inferred changes in the light requirements based , on changes in standing crop (Congdon and McComb 1979). Ott (1979) presented evidence foranannualrhythminthegrowthcycleofPosidoniaoceanica. Aftertwoyearsofgrowth in vivo under conditions of constant light and P. oceanica still temperature, the seagrass exhibited a seasonal pronounced growth rhythm. However, changes in photosynthetic capabilities were not examined in this study. Bulthuis (1983b) reported that temperature dramaticallyinfluencedthePvs.IcurveofHeterozosteratasmanica, Bulthuissuggestedthat Heterozostera has higher light requirements during summer than during winter due to increased respiration associated with high summer temperatures. Thus, this species exhibits It was seasonal changes in its light requirements due to changes in the rate of respiration. also suggested that these plants are more susceptible to decreased light availability during summer seasonal than during winter (Bulthuis 1983b). Macauley et al (1988) examined changes in standing crop and chlorophyll content of Thalassia testudinum in the Northern Gulf of Mexico. Thalassia exhibited seasonal variation in standing crop and chlorophyll content; standing crop was strongly correlated with temperature and moderately correlated with incident irradiance. They suggested that temperature was the controlling factor for ThalassiaproductivityalongtheNorthernGulfofMexico(Macauleyetal, 1988).Perezand Romero (1992) found that Cymodocea nodosa exhibits some photoadaptation to the yearly and P lightcycle. Duringsummer,whenthereishighirradiance,theplantshavehighI k max values and do not exhibit photoinhibition; however, during the winter these plants shift towardshade-adaptedcharacters(Table 12).Itissuggestedthatwinterphotosyntheticrates may be limited by low temperatures rather than low light. The mechanisms causing these adaptations are not well understood and may involve short-term light adaptation and internal rhythms keyed to changes of the seasons. Researchers have also investigated seasonal changes in the organic constituents of seagrasses. Dawes and Lawrence (1979) investigated how blade removal influenced the proximate composition of the rhizome of Thalassia testudinum. They concluded that the rhizome acts as a storage organ which supports blade regeneration as well as seasonal growth. Soluble carbohydrate was the primary reserve and was actively mobilized through the rhizome. Rhizomes seasonal rise had the largest carbohydrate reserves exhibiting a Table 12. Seasonal variation in the P-I parameters of Cymodocea nodosa. P and R max expressed in mg 02 per d dry h'1; I and expressed as fimo\ m'2 s'1 From Perez and Mt . Romero 1992. Month P R max February 2.31 0.41 69 10 April 2.64 0.14 238 12 June 5.48 0.47 398 31 August 4.37 0.56 88 10 November 2.51 0.73 160 36 December 3.22 1.00 127 30 during late spring and summer and a decrease during winter. Carbohydrate levels in the Dawes and Lawrence made seasonal rhizome were lowest in early spring. (1980) a examination of the proximate constituents of Thalassia testudinum Halodule wrightii and , Syringodium filiforme (Table 13). They found that in the rhizomes soluble carbohydrate levels were highest during the fall and lowest during spring. They suggest that soluble low. Calorific carbohydrates sustained the plants through the winter when productivity was levels were similar between species with the highest levels occurring in the rhizomes. Of the three species, the rhizomes of Thalassia had the highest level of organic material, be more tolerant of adverse conditions. Dawes and suggesting that this species may Lawrence (1983) suggest that Thalassia is better adapted to year round growth, while the other species (Halodule and Syringodium) are more opportunistic. Pirc(1985)examinedthegrowthdynamicsofPosidoniaoceanicawithrespect tothe seasonal changes of soluble carbohydrates, starch and other organic compounds in different parts of the plant. He concluded that winter leaf growth was supported by the mobilization of starch from the rhizomes. Pirc (1985) hypothesized that subsidized winter leaf growth enabled the plant to take maximal advantage of increased light during the spring. During summer and autumn large quantities of carbohydrates found in both the leaves and were the rhizomes. Plants utilize summer and fall influx to energy compensate for lowproduction during winter and spring. Ralph et al. (1992) examined the distribution of extractable carbohydrate reserves in the rhizome of Posidonia australis. Carbohydrate levels were reserves were significantly higher in the stele tissue than in the cortex. Lower carbohydrate near found in juvenile tissue and the apical meristem, while unexpanded internodes had Table 13. Seasonal variation of protein, soluble carbohydrate and kilocalories for blades From Dawes 1987. and rhizomes of three seagrass species. January April July October Thalassia testudinum Blades Protein 8 9 22 13 Carbohydrate 6 9 9 7 Kilocalories 2.4 3.0 3.1 2.6 Rhizome Protein 9 8 16 7 Carbohydrate 12 21 24 36 Kilocalories 3.2 3.4 3.0 2.8 v Syringodium filiforme Blades Protein 9 8 13 13 Carbohydrate 22 16 18 20 Kilocalories 3.1 2.4 3.2 3.1 Rhozome Protein 9 5 12 16 Carbohydrate 36 38 50 46 Kilocalories 3.6 3.7 3.6 3.5 Halodule wrightii Blades Protein 19 18 19 14 1419 15 13 Carbohydrate Kilocalories 3.1 3.5 3.3 3.3 Rhizome Protein 9 7 8 8 Carbohydrate 43 40 43 54 Kilocalories 3.7 3.7 3.4 3.6 Percent dry weight 2Per gram dry weight It was are used relatively large carbohydrate reserves. suggested that stored carbohydrates for winter maintenance and support early spring blade growth. Seagrasses can also adjust the concentration and distribution of chlorophyll within their tissues. Mazzella et aL (1979) concluded that the photosynthetic activity of Zostera marina leaves was and regulated by four factors: age of tissue, light intensity, exposure For of epiphytes. young parts of leaves, light intensity and tissue maturity were presence most important. For older leaves, the presence of epiphytes was more important than the other factors. Wiginton and McMillan (1979) examined the chlorophyll composition of seagrasses under controlled light conditions. Chlorophyll concentrations in both Thalassia and Halodule were correlated with chlorophyll content increased with decreasing light in both the field and lab. Chlorophyll a:b ratios were correlated with different depth ranges and affect depth distribution of seagrasses. The similarities in light levels at the may in both Croix and maximum depth of seagrasses St. Texas suggests that the seagrass were restricted at each locale. In populations by similar light relationships general, are seagrasses respondingmainlytochangesinlightquantity,notlightquality(Wigintonand McMillan 1979). Mazzella et al (1980) found that Zostera had a gradient of leaf pigmentation and photosynthetic activity. The initial slope of the P vs. I curve increased fromleafbasetoleaftip. Lightsaturationoccurredat100-150/nmolm2s'1forallleaf types. Czeczuga (1986) examined the effect of light quality on the photosynthetic pigments of the green alga Chora. In freshwater Chora forms meadows that have the same may functions as seagrasses in estuarine waters. Highest chlorophyll a,b and carotenoids were found when green and yellow filters were placed over the light source, while the lowest pigmentconcentrations occurredwhenblueandredfilterswereused. Enriquezetal(1992) examined how light was absorbed byPosidonia oceanica. Their results indicate the amount of light absorbed increases linearly with increased pigment packaging in the leaves. Thus, in increased The increasing chlorophyll results light absorption per unit leaf weight. increasing absorption per unit weight should increase the photosynthetic and growth rates of a light-limited plant. in the concentration allow a to survive in a Changes chlorophyll may species particularhabitat. Severalstudieshaveinvestigatedthephysiologicalecologyofseagrasses in relation to Dennison et light and photosynthesis. aL (1981) concluded, based on shading and reflecting studies, that Zostera marina adjusts to light conditions by changes in leaf area production. Under low light the plants increase their leaf area to intercept more photons. Jimenez et aL (1987) examined the of Zostera noltii and Z. marina to high light response under stressed conditions. Zostera marina had higher chlorophyll levels at low light, with maximumchlorophyllconcentrationsoccurringat150pmolm'2 1 Zosteranoltiichlorophyll s’ . increased with light intensity and were constant at light levels above 150 pmol m’2 s’1 . 21 Zostera noltii was light saturated at 3600 pmol m’ s’ showing no evidence of photoinhibition,while Z.marinawaslightsaturatedat1100nmolm’2s’1withsignificant was photoinhibition above this light intensity. For both speciesI 30-35 /imol m’2 s’1 They c. concluded that Z. noltii was more photosynthetically efficient, especially at high irradiances. Dawes et al (1989) compared the physiological ecology of Halophila dedpiens and H. from Florida. johnsonii With respect to P vs. I characters, H. dedpiens was strongly photoinhibitedatirradiances above 300 /xmolm’2s’1while H.johnsonii wasphotoinhibited 21 2 atirradiancesabove 1000pmolm' s'HalophiladecipienshadanI valueof29pmolm' .c 1 21 s'compared to about 50 /imol m' s' for H. johnsonii. Although interannual variation may be large, most seagrass species exhibit very pronounced seasonal trends. In general, plant biomass increases rapidly during spring and earlysummerandismaintainedthroughthesummer. Duringfall,biomass decreasesasthe blades senesce and slough off of the plant. The P vs. I characters of most seagrass species followsimilartrend. and "shade a The plants appear to be "sun adapted" during the summer adapted" during the winter. Generally, P is lower during the winter than during the max The summer; presumably, this is due to reduced enzymatic activity at lower temperatures. concentration of organic constituents also varies seasonally. During summer, when production is greatest, the plants store large amounts of carbohydrate in the rhizomes. During the fall and winter, when light levels are low, the plants draw upon these reserves. During late winter (very early spring) the plants put forth a new crop of leaves, supported at least in part by stored carbohydrate. During this time the plants are probably most susceptible to prolonged periods of reduced irradiance. CHAPTER 5: RECOMMENDATIONS FOR FURTHER RESEARCH Texas estuarine systems are remarkably diverse, as examplified from the hypersaline Laguna Madre to the freshwater dominated estuaries of the northern Texas coast. As a to water consequence, the potential problems faced by resource managers with respect quality and water clarity (transparency) is unique to each estuary. Estuarine systems dominated by freshwater inflow are more likely to experience the effects of eutrophication than estuaries where evaporation exceeds inflow (e.g., Laguna Madre). Conversely, systems that are are flushed frequently less likely to experience the long-term effects of chronic phytoplankton blooms, as is now occurring in Laguna Madre (the brown tide). Setting water quality criteria to maintain the current level of productivity in these systems will therefore be estuary specific. The immediate challenge is to thoroughly define hydrographic characteristics common to each before to to system they begin respond negatively anthropogenic inputs. The goal will then be to use this database to establish water quality standards to maintain the health of each system. With very few exceptions, Texas is fundamentally different than most other East Coast and Gulf Coast states which have lost much of their valuable coastal habitat; in Texas the goal is still to preserve the natural habitat, not to restore or create what has been lost. The extensive development ofsubmerged aquatic vegetation in south Texas bays and estuaries is primarily related to reduced freshwater inflows and lower levels of inorganic nutrients. light attenuation is the major problem that must be addressed in these systems. Although increasing urban and agricultural development do present an alarming potential for eutrophication, dredging and other construction activities clearly present the greatest immediate danger to seagrass beds. These activities significantly increase turbidity which reduces the amountof light reaching the bottom. of beds now has Our failure to largely prevent the loss seagrass and in the past largely resulted from an inadequate knowledge of: (1) the minimum light requirements needed to maintain a positive carbon balance and net growth for the various species, (2) the amount of light actually reaching the plants on the seabed. Although some effort has been made to address the photosynthetic light these requirements of seagrasses through laboratory experiments, measurements often cannot be extrapolated into the field. In situ measurements using entire plants thus are highly recommended and potentially the most useful for management purposes. Currently, we have information on the light requirements and underwater light fields foronlyoneoffiveTexasseagrasses (Halodulewrightii). Insituphotosyntheticparameters need to be measured for both Syringodium filiforme and Thalassia testudinum which along , with Halodule of meadows in constitute the majority of the approximate 850 km2 , seagrass Texas. Without the knowledge of the minimum light requirements for growth and photosynthesis of these plants, the development of water quality and water transparency standards will be extremely difficult. Compilation of published nutrient and water chemistry data for each estuary with information on the in situ light requirements and underwater light fields ofTexas seagrasses can be used to develop water quality criteria to in Texas estuaries. The preserve seagrasses contributions of suspended solids, chlorophyll a and dissolved inorganic nitrogen to the diffuse light attenuation coefficient (k) for each estuary can then be used in carbon budget models for seagrasses within each estuary to maintain the current productivity and distribution ofseagrasses. This approach is similar to that used in Chesapeake Bay (Batuik et al 1992; Dennison et al 1993) but incorporates quantitative data on underwater light ., ., fields and the light requirements of the plants. Recent studies suggest that the physiological Halodule is similar estuaries light requirements of the seagrass wrightii in Texas among (Dunton, submitted and unpub. data); if this is true for other species then the strategy described above is not unnecessarily complex, and our goal of establishing water quality standards in Texas coastal environments is attainable. to protect seagrasses Based on the arguments presented above, our recommendations for further work to achieve this goal are: 1. Determine the in situ photosynthetic requirements ofTexas seagrasses, including (by priority): Thalassia testudinum, Syringodium filiforme, Halophila engelmanni and Ruppia maritima (this work on Halodule wrightii is complete). 2. CollectinsitucontinuousmeasurementsofunderwaterPARinconjunctionwith nutrient and chlorophyll measurements in Texas estuaries containing seagrasses for calculation of diffuse light attenuation coefficients (k). (This work has been in progress in both Upper Laguna Madre and Corpus Christi Bay since 1990/1991; see Dunton, submitted). 3. Evaluate the contributions of suspended solids, water column chlorophyll a, inorganic nitrogen and dissolved matter to the diffuse light attenuation coefficient (k) for specific Texas estuaries based on available data and field sampling. 4, Develop seasonal and annual carbon budgets for predominant Texas seagrasses. This information can be used to: (a) assess the impacts of chronic light reduction, (b) minimize loss of habitat by temporal or seasonal restriction of construction activities and, (c) determine the feasibility of mitigation (creation or restoration) projects. Literature Cited Ascherson, P. and P. Graebner. 1907. Potamogetonaceae. Pflanzenreich Heft 31:1-184, f. 1-36. (as in den Hartog 1970). Backman, T.W. and D.C. Barlotii. 1976. Irradiance reduction: effects on standing crop of Mar. Biol. 34: 33-40. the eelgrass Zostera marina in a coastal lagoon. Barnabas, A.D. 1989. Apoplastic tracer studies in the leaves of a seagrass. 11. Pathway into leaf veins. Aquat. Bot. 35: 375-386. Barnabas, A.D. 1991. Thalassodendron ciliatum (Forssk.) den Hartog: root structure and 129-143. histochemistry in relation to apoplastic transport. Aquat. Bot. 40: Barnabas, A.D. and H.J. Arnott. 1987. Zostera capensis Setchell: root structure in relation to function. Aquat. Bot. 27: 309-322. Batuik, R.A., R.J. Orth, K.A. Moore, W.C. Dennison, J.C. Stevenson, L.W. Staver, V. Carter, N.B. Rybicki, R.E. Hickman, S. Kollar, S. Bieber, P. Heasley. 1992. Chesapeake Bay submerged aquatic vegetation habitat requirements and restoration targets: a technical synthesis. U.S. Environmental Protection Agency, Annapolis Maryland. 186 pp. Bellrose,F. 1976. Ducks,geeseandswansofNorthAmerica. StackpoleBooks,Harrisburg Pa. 540 pp. Benz, M.C., NJ. Eiseman and E.E. Gallaher. 1979. Seasonal occurance and variation in standing crop of a drift algal community in the Indian River, Florida. Bot. Mar. 22: 413-420. 1983. Borum, J. The quantitative role of macrophytes, epiphytes, and phytoplankton under different nutrient conditions in Roskilde Fjord, Denmark. Proceedings of the International Symposium on Aquatic Macrophytes. Faculty of Science, Nijmejen, The Netherlands. 35-40. Bothwell, M.L. 1989. Phosphorus-limited growth dynamics of lotic periphytic diatom communities: areal biomass and cellular growth rate Can. J. Fish. Aquat. responses. Sci. 46: 1293-1301. Bougis, P. 1976. Marine plankton ecology. North-Holland Publishing Co., Amsterdam. 355 pp. Boysen-Jensen, P. 1914. Studies concerning the organic matter of the sea bottom. Rep. Danish Biol. Stn. 22: 1-39. 1974. on Turtle Buesa, RJ. Population and biological data grass (Thalassia testudinum Konig, 1805) on the Northwestern Cuban Shelf. Aquaculture 4: 207-226. Bulthuis, D.A. 1983a. Effects of in situ light reduction on density and growth of the Heterozostera tasmanica (Martens ex Aschers.) den Hartog in Western Port, seagrass Victoria, Australia. J. Exp. Mar. Biol. Ecol. 67: 91-103. Bulthuis, D.A. 1983b. Effects of temperature on the photosynthesis-irradiance curve of the Australian Heterozostera tasmanica. Mar. Biol. Lett. 4: 47-57. seagrass, 1987. Bulthuis, D.A. Effects of temperature on photosynthesis and growth of seagrasses. Aquat. Bot. 27: 27-40. 1984. Bulthuis, D.A., G.W. Brand and M.C. Mobley. Suspended sediments and nutrients in water ebbing from seagrass-covered and denuded tidal mudflats in a Southern Australian embayment. Aquat. Bot. 20: 257-266. Burch, T.A. 1983. Inventory of submerged vegetation in Choctawhatchee Bay, Florida. Water Resources Northwest Florida Water Management District. Special Report 93-4. 25 pp. Burkholder, P.R. and T.E. Doheny. 1968. The Biology of eelgrass. Dept, of Conservation and Waterways, Hempstead, N.Y. 120 pp. Burrell, D.C. and J.R. Schubel. 1977. Seagrass ecosystem oceanography. In: C.P. Mcßoy andC.Helfferich(eds).Seagrassecosystems: ascientificperspective.MarcelDekker, Inc. New York. 196-232 pp. Buskey, EJ. and D.A. Stockwell. 1993. Effects of a persistent "Brown Tide" on zooplanktonpopulationsintheLagunaMadreofSouthTexas.In: T.J.Smaydaand Elsevier Science Publ. Y. Shimizu (eds.) Toxic phytoplankton blooms in the sea. 659-666 pp. 1986. The loss Cambridge, ML., A.W. Chiffings, C. Brittan, L. Moore and AJ. McComb. of in Cockbum Sound, Western Australia. 11. Possible causes of seagrass seagrass decline. Aquat. Bot. 24: 269-285. 1984. The loss of in Cockbum Cambridge, M.L. and AJ. McComb. seagrasses Sound, Western Australia. I. The time course and magnitude ofseagrass decline in relation to industrial development. Aquat. Bot. 20: 229-243. 1985. The effects of the Carter, V. and N.B. Rybicki. grazers and light penetration on survival of transplants of Vallisneria americana Michx in the tidal Potomac River, Maryland. Aquat. Bot. 23: 197-213. Carter, V. and-N.B. Rybicki. 1990. Light attenuation and submersed macrophyte distribution in the tidal Potomac River and Estuary. Estuaries 13(4): 441-452. Chambers, P.A and J. Kalff. 1985. Depth distribution and biomass of submersed aquatic macrophyte communities in relation to secchi depth. Can. J, Fish. Aquat. Sci. 42: 701-709. Chesher, R.H. 1971. Biological impact ofa large-scale desalination plant at Key West. U.S. Environmental Protection Agency, Washington, D.C. Water Pollution Control Research Series 18080 GBX 12/71. 150 pp. Christiansen, C, H. Christoffersen, J. Dalagaard and P. Nornberg. 1981. Coastal and near- shore changes correlated with the die-back in eelgrass (Zostera marina L.). Sed. Geol. 28:163-173. Congdon, R.A. and A.J. McComb. 1979. Productivity of Ruppia: seasonal changes and 121-132. dependence on light in an Australian estuary. Aquat. Bot. 6: 1988. Cooper, L.W. and C.P. Mcßoy. Anatomical adaptations to rocky substrates and surf exposure by the seagrass genus Phyllospadix. Aquat. Bot. 32: 365-381. Cornelius, S.E. 1977. Food and resource utilization by winteringredheads on lower Laguna Madre. J. Wildl. Manag. 41(3): 374-385 Cosper, E.M., W.C. Dennison, EJ. Carpenter, V.M. Bricelj, J.G. Mitchell, S.H. Kuenstner, D. Colflesh and M. Dewey. 1987. Recurrent and persistent brown tide blooms perturb coastal marine ecosystem. Estuaries 10(4): 284-290. Costa, J.E. 1988. Eelgrass in Buzzards Bay: distribution, production and historical changes in abundance. EPA BBP-88-05. 204 pp. 1978. Cowper, S.W. The drift algal community of seagrass beds in Redfish Bay, Texas. Contrib. Mar. Sci. 21: 125-132. Cox, P.A. 1988. Hydrophilous pollination. Ann. Rev. Ecol. Syst. 19: 261-280. 1981. An integrated system of classification of flowering plants. Columbia Cronquist, A. University Press. N.Y. on the content Czeczuga, B. 1986. The effect of light of photosyntheticlly active pigments in species of the genus Chora. Aquat. Bot. 24: 397-401. 1987. oftheGulfofMexicoandFloridaCoasts.In: Dawes, C.J. The dynamic seagrasses MJ. Durako, R.C. Phillips and R.R. Lewis (eds.). Proceedings of the symposium on ofthe southeastern United States. Florida Mar. Res. subtropical-tropical seagrasses Publ., Florida Dept. Nat. Res. Number 42. 25-38 pp. Dawes, C.J. and J.M. Lawerence. 1979. Effects of blade removal on the proximate Thalassia testudinum Banks composition of the rhizome of the seagrass ex Konig. Aquat. Bot. 7: 255-266. Dawes, CJ. and J.M. Lawerence. 1980. Seasonal changes in the proximate constituents of the Thalassia testudinum Halodule seagrasses , wrightii, and Syringodium filiforme. Aquat. Bot. 8: 371-380. J.M. Lawerence. 1983. Dawes, CJ. and Proximate composition and caloric content of Mar. Tech. Soc. J. 17(2): 53-58 seagrasses. Dawes, CJ., C.S. Lobban and D.A. Tomasko. 1989. A comparison of the physiological ecologyofthe seagrassesHalophila ostenfeldandH. JohnsoniiEisemanfromFlorida. Aquat. Bot. 33: 149-154. Dawes, CJ. and D.A. Tomasko. 1988. Depth distribution of Thalassia testudinum in two meadows on the West coast of Florida; a difference in effect of light availability. P.5.Z.N.1.: Mar. Ecol. 9(2): 123-130. of the world. North-Holland den Hartog, C. 1970. Seagrasses Publishing Comp Amsterdam. 275 pp. Foreward. den Hartog, C. 1980. In: R.C. Phillips and C.P. Mcßoy (eds). Handbook of seagrass biology: an ecosystem perspective. Garland STPM Press NY. ix-xiii pp. Dennison, W.C. 1987. Effects of light on seagrass photosynthesis, growth and depth distribution. Aquat. Bot. 27:15-26. Dennison, W.C. and R.S. Alberte. 1982. Photosynthetic responses of Zostera marina L. (eelgrass) to in situ manipulations of light intensity. Oecologia 55:137-144. Dennison,W.C.andR.S.Alberte. 1985.Roleofdailylightperiodinthedepthdistribution of Zostera marina (eelgrass). Mar. Ecol. Prog. Ser. 25:51-61. 1986. Dennison,W.C.andR.S.Alberte. PhotoadaptationandgrowthofZosteramarinaL. (eelgrass) transplants along a depth gradient. J. Exp. Mar. Biol. Ecol. 98:265-282 1989. Dennison, W.C., GJ. Marshall and C. Wigand. Effect of "brown tide" shading on eelgrass (Zostera marina L.) distributions. In: E.M. Cosper, V.M. Bricelj and E.J. Coastal and Estuarine Studies No. Carpenter (eds). Novel phytoplankton blooms. 35. Springer-Verlag, Berlin. 675-692 pp. 1981. Dennison, W.C., D. Mauzerall and R.S. Alberte. Photosynthetic response of Zostera marina (eelgrass) to in situ manipulations of light. Biol. Bull. 161(2):311-312. Dennison, W.C., RJ. Orth, K.A. Moore, J.C. Stevenson, V. Carter, S. Kollar, P.W. 1993. Bergstrom and R.A. Batuik. Assessing water quality with submersed aquatic vegetation. BioScience 43(2): 86-94. Drew, E.A. 1978. Factors affecting photosynthesis and its seasonal variation in the seagrasses Cymodocea nodosa (Ucria) Aschers, and Posidonia oceanica (L.) Delile in the mediterranean. J. Exp. Mar. Biol. Ecol. 31: 173-194. Drew, E.A. 1979. Physiological aspects of primary production in seagrasses. Aquat. Bot. 7: 139-150. Duarte, C.M. 1991. Seagrass depth limits. Aquat. Bot. 40: 363-377. Dunton, K.H. 1990. Production ecology of Ruppia maritima L. s.l. and Halodule wrightii Aschers. in two subtropical estuaries. J. Exp. Mar. Biol. Ecol. 143: 147-164. Dunton, K.H. submitted. An Annual quantum budget for the subtropical seagrass Halodule wrightii. Mar. Biol. Dunton, K.H. and D.A. Tomasko. submitted. In situ photosynthetic performance in the seagrass Halodule wrightii in a hypersaline subtropical lagoon. Mar. Ecol. Prog. Ser. Durako, MJ. and M.O. Hall. 1992. Effects of light on the stable carbon isotope composition of the seagrass Thalassia testudinwn. Mar. Ecol. Prog. Ser. 86:99-101. 1987. Eckman, J.E. The role of hydrodynamics in recruitment, growth, and survival of Argopecten irradians (L.) and Anomia simplex (D’Orbigny) within eelgrass meadows. J. Exp. Mar. Biol. Ecol. 106: 165-191. Eleuterius, L.N. 1989. Catastrophic loss of seagrasses in Mississippi Sound. MMS-ITM Abstract. New Orleans. 1992. Posidonia Enriquez, S., S. Agusti and C.M. Duarte. Light absorption by seagrass oceanica leaves. Mar. Ecol. Prog. Ser. 86: 201-204 Evans, A.S. 1984. Temperature adaptation in seagrasses. Masters thesis. College of William and Mary, Virginia. 75 pp. Fenchel, T. 1977. Aspects of the decomposition of seagrasses. In: Mcßoy, C.P. and C. Helfferich(eds).Seagrassecosystems: ascientificperspective.MarcellDekker Inc., N.Y. 123-145 pp. Fonseca, M.S. and J.A. Cahalan. 1992. A preliminary evaluation of wave attenuation by Est. Coast. Shelf Sci. 35: 565-576. four species of seagrass. 1982. Influence ofthe Fonseca, M.S., J.S. Fisher, J.C. Zieman and G.W. Thayer. seagrass, Zostera marina L., on current flow. Est. Coast. Shelf Sci. 15:351-364. J.C. Zieman. 1991a. Fourqurean, J.W. and Photosynthesis, respiration and whole plant carbonbudgetofthe Thalassiatestudinum.Mar.Ecol.Prog.Ser.69:161-170 seagrass J.C Zieman. 1991b. Fourqurean, J.W. and Photosynthesis, respiration and whole plant carbon budgets of Thalassia testudinum, Halodule wrightii and Syringodium filiforme. In: Kenworthy, WJ. and D.E. Haunert (eds.). The light requirements ofseagrasses: proceedings of a workshop to examine the capability of water quality criteria, standards and to NOAA Technical monitoring programs protect seagrasses. Memorandum NMFS-SEFC-287. 59-70 pp. 1984. Fry, B. 13C/12C ratios and the trophic importance of algae in Florida Syringodium filiforme seagrass meadows. Mar. Biol. 79: 11-19. Gessner, F. 1961. Hydrostatischer druk und pflanzenwachstum. In: W. Ruhland (ed.) Encyclopedia of plant physiology. Vol. 16, Springer-Verlag, Berlin, 668-690 pp. Giesen, W.BJ.T., M.M. van Katwijk and C. den Hartog. 1990. Eelgrass condition and turbidity in the Dutch Wadden Sea. Aquat. Bot. 37: 71-85. K.B. Clark. 1981. Seasonal variation of the Gilbert, S. and in standing crop seagrass Syringodium filiforme and associated macrophytes in the Northern Indian River, Florida. Estuaries 4(3): 223-225. Ginsburg, R.N. and H.A. Lowenstam. 1958. The influence of marine bottom communities J. Geol. 66: 310-318. on the depositional environment of sediments. 1988. of Goldsborough, WJ. and W.M. Kemp. Light responses a submersed macrophyte: 1775-1786. implications for survival in turbid tidal waters. Ecology 69(6): Goodwin, H. and L. Goodwin. 1976. The Indian River-an American lagoon. Compass Publications. Arlington, Va. 66 pp. 1983. Fisheries habitat loss in Harris, 8.H., K.D. Haddad, K.A. Steidinger and J.A. Huff. Charlotte Harbo and Lake Worth, Florida. Final Rep. to the Fla. Dept. 211 Environmental Regulation. Coastal Zone Management Grant CM-29. pp. 1976. structure and the effects Hech, K.L., Jr. Community of pollution in seagrass meadows and adjacent habitats. Mar. Biol. 35:345-357. Hogan, W.A, 1983. Computer simulation of light attenuation. Florida Sci. 46(3/4): 145­ 150. 1976. An inshore marine invertebrate Hooks, T.A., L. Heck, Jr., and RJ. Livingston. structure and habitat associations in the northeastern Gulf of Mexico. community: Bull. Mar. Sci. 26:99-109. Hoven, H.M. 1992. Eelgrass {Zostera marina L.) as a substratum for larval blue mussels (Mytilus edulis L.) in Mount Desert Narrows, Maine. Masters Thesis. University of New Hampshire. Durham, New Hampshire. 100 pp. Hutchinson, I. 1934. The families of flowering plants. 11. Monocotyledons. Ed. 1. London, xiii+243 pp. (as in den Hartog 1970). Iverson, R.L. and H.F. Bittaker. 1986. Seagrass distribution and abundance in Eastern Gulf of Mexico coastal waters. Est. Coast, and Shelf Sci. 22: 577-602. 1987. Jim6nez, C., F.X. Niell and P. Algarra. Photosynthetic adaptation of Zostera noltii Hornem. Aquat. Bot. 29: 217-226. 1992. Kaldy, J.E. Plant community response to the simulated eutrophication of eelgrass (Zostera marina L.) beds using experimental mesocosms. Masters Thesis. University 76 of New Hampshire, Durham, New Hampshire. pp. Short, A.C. Mathieson and D.M. Burdick, in The of Kaldy, J.E., F.T. prep. process eutrophication in habitats dominated by eelgrass (Zostera marina L.): light and nutrient manipulations restructure the eelgrass community. Manuscript in prep. 1991. Kautsky,H. Influenceofeutrophicationonthedistributionofphytobenthicplantand animal communities. Int. Revue ges. Hydrobiol. 76(3): 423-432. 1991. Kenworthy, WJ. and D.E. Haunert (eds.). The light requirements of seagrasses: proceedings of a workshop to examine the capability of water quality criteria, standards and to Technical monitoring programs protect seagrasses. NOAA Memorandum NMFS-SEFC-287, 181 pp. Kerr, E.A. and S. Strother. 1985. Effects of irradiance, temeperature and salinity on 23: 177-183. photosynthesis of Zostera muelleri. Aquat. Bot. 1984. Kitting, C.L., B. Fry and M.D. Morgan. Detection of inconspicuous epiphytic algae meadows. 145-149. supporting food webs in seagrass Oecologia 62: Kirk, J.T.O. 1983. Light and photosynthesis in aquatic ecosystems. Cambridge University 401 Press, Cambridge. pp. Kirkman, H. 1978. Decline of seagrass in northern areas of Moreton Bay, Queensland. Aquat. Bot. 5: 63-76. 1981. Kulczycki, G.R., R.W. Virnstein and W.G. Nelson. The relationship between fish Est. Coast. Shelf abundance and algal biomass in a seagrass-drift algae community. Sci. 12: 341-347. 1980. Lewis, R.R, 111, and R.C. Phillips. Seagrass mapping project. Hillsborough County, Florida. 15 Report No. 1, Tampa Bay Cooperative Seagrass Project. pp. 1985. of Lewis, R.R, M.J. Durako, M.D. Moffler, and R.C. Phillips. Seagass meadows Tampa Bay-A review. Pp. 210-246 in S.F. Treat, J.L. Simon, R.R. Lewis, HI, and R.L. Whitman, Jr., (eds). Prodeedings, Tampa Bay Area Scientific Information Symposium. Florida Sea Grant Publ. 65. Burgess Publishing Co., Minneapolis. M. 1986. of Posidonia oceanica and its Libes, Productivity-Irradiance relationship epiphytes. Aquat. Bot. 26: 285-306. 1972. The Livingston, RJ., T.S. Hopkins, J.K. Adams, M.D. Schmitt, and L.M. Welch. effects of dredging and eutrophication Mulat-Mulatto Bayou (Escambia Bay; on Pensacola, Florida). Unpublished report, Florida Department of Transportation. Livingston, RJ. 1975. Impact of kraft pulp-mill effluents on estuarine and coastal fishes Mar. Biol. 32:19-48. in Apalachee Bay, Florida, U.S.A. Environmental Livingston, R.J. 1979. impact assessment regarding the drilling of an exploratory oil well in East Bay of the Pensacola Bay system (Santa Rosa County, Florida). Unpublished report. RJ. 1980c. Critical habitat assessment of the and Livingston, Apalachicola estuary associated coastal areas. Unpublished report. 1982a. coastal Mar. Livingston, RJ. Trophic organization of fishes in a seagrass system. Ecol. Prog. Ser. 7:1-12. 1983. Resource Florida Sea Grant Livingston, RJ. atlas of the Apalachicola estuary. Report Number 55. 64 pp. 1984a. Livingston, RJ. Trophic response of fishes to habitat variability in coastal seagrass systems. Ecology 65:1258-1275. Livingston, RJ. 1987. Historic trends of human impacts on seagrass meadows in Florida. In:M.J.Durako,R.C.PhillipsandR.R.Lewis(eds.). Proceedingsofthesymposium on subtropical-tropical seagrasses of the Southeastern United States. Florida Dept. Nat. Resources, Bureau of Marine Research, Florida Marine Research Publications number 42. 139-151 pp. Joining, K. 1990. Seaweeds. John Wiley & Sons Inc., N.Y. 527 pp. Macauley, J.M., J.R. Clark and W.A. Price. 1988. Seasonal changes in the stading crop and chlorophyll content of Thalassia testudinum Banks ex Konig and its epiphytes in the Northern Gulf of Mexico. Aquat. Bot. 31: 277-287. Madsen, J.D., and M.S. Adams. 1989. The light and temperature dependence of photosysnthesis and respiration in Potamogeton pectinatus L. Aquat. Bot. 36: 23-31. 1936. Blutenbau und Verwandtschaft bei den einfashsten Helobidae. Ber. Markgraf, F. Deutsch. Bot. Ges. 54: 191-229. (as in den Hartog 1970). Marsh, J.A., W.C. Dennison and R.S. Alberte. 1986. Effects of temperature on photosynthesis and respiration in eelgrass (Zostera marina L.). J. Exp. Mar. Biol. Ecol. 101: 257-267. Mauzerall and R.S. Alberte. 1979. characteristics of Mazzella, L., D. Photosynthetic Zostera marina L. (eelgrass). Biol. Bull. 157(2): 382. D. Mauzerall and R.S. Alberte. Mazzella, L,1980. Photosynthetic light adaptation features of Zostera marina L. (eelgrass). Biol. Bull. 159(2): 500. 1970. J. Wildl. McMahan, C.A. Food habits of ducks wintering on Laguna Madre, Texas. Manag. 34(4): 946-949. 1961. effects McNulty, J.K. Ecological of sewage pollution in Biscayne Bay, Florida: sediments and distribution of benthic and fouling organisms. Bull. Mar. Sci. Gulf Carib. 11:394-447. McNulty, J.K., W.N. Lindall, Jr., and J.E. Sykes. 1972. Cooperative Gulf of Mexico I: area Natl. Ocean. estuarine inventory and study, Florida, phase description. Atmos. Admin. Tech. Rep. Int. Mar. Fish. Serv. Circ. 368:1-126. Megard, R.O. and T. Berman. 1989. Effects of algae on the Secchi transparency of the southeastern Mediterranean Sea. Limnol. Oceanogr. 34(8): 1640-1655. Merkord, G.W. 1978. The distribution and abundance of seagrasses in Laguna Madre of Texas. Masters thesis. Texas A&I University, Kingsville Texas. 56 pp. Meulemans, J.T. 1987. A method for measuring selective light attenuation within a periphytic community. Arch. Hydrobiol. 109(1): 139-145. Milne, LJ. and M.J. Milne. 1951. The eelgrass catastrophe. Sci. Amer. 184: 52-55. Muehlstein, L.K., D.Porter and F.T. Short. 1991. Labyrinthula zosterae Sp. Nov., the causativeagentofwastingdiseaseofeelgrass,Zosteramarina.Mycologia83(2): ISO­ -191. National Oceanic and Atmospheric Administration. 1988. Galveston Bay Seminar: executive summary. U.S. Department of Commerce. Washington, DC: US Government Printing Office. 28 pp. National Oceanic and Administration. 1990a. Atmospheric Fifty years of population the Nations Coasts 1960-2010. US of Commerce. change along Department Washington, DC: U.S. Government Printing Office. 41 pp. National Oceanic and Atmospheric Administration. 1990b. Estuaries ofthe United States. Vital statistics of a National Resource Base. US Department of Commerce. Washington, DC: U.S. Government Printing Office. 79 pp. Neverauskas, V.P. 1988. Response of a Posidonia community to prolonged reduction in light. Aquat. Bot. 31:361-366. Odum, H.T. 1963. Productivity measurements in Texas turtle and the effects of grass dredging an intracoastal channel. Pub, Instit. Mar. Sci. 9: 48-58. Odum, M.A 1985. Shading effects of epiphytes on leaves of turtle grass Thalassia testudinum Banks ex Konig. Masters Thesis. The University of Texas at Austin, Austin, TX. 79 pp. Olinger, L.W., R.G. Rogers, P.L. Fore, R.L. Todd, B.L. Mullins, F.T. Bisterfeld, and L.A. 11. 1975. Environmental and studies of Escambia and the Wise, recovery Bay Pensacola Bay system, Florida. U.S. Environmental Protection Agency, Region IV. Atlanta, Ga. Onuf, C.P. 1993. Seagrasses, dredging and light in Laguna Madre, Texas, USA Est. Coast. Shelf Sci. In press. S. and R. Benner. 1993. Opsahl, Decomposition of senescent blades of the seagrass Halodule wightii in a subtropical lagoon. Mar. Ecol. Prog. Ser. 94:191-205. Orth, R.J., K.L. Heck and J. van Montfrans. 1984. Faunal communities in beds: seagrass areviewoftheinfluenceofplantstructureandpreycharacteristics onpredator-prey relationships. Estuaries 7(4a): 339-350. Orth, RJ. and K.A. Moore. 1984. Distribution and abundance of submerged aquatic vegetation in Chesapeake Bay: an historical perspective. Estuaries 7(4b): 531-540. Orth, R.J. and K.A. Moore. 1988. Distribution ofZostera marina L. and Ruppia maritima L. sensu lato along depth gradients in the lower Chesapeake Bay, U.S.A. Aquat. Bot. 32: 291-305. 1964. The Flowers, Fruits, and seeds of Thalassia testudinum Orpurt, P.A. and L.L. Boral. Koenig. Bull. Mar. Sci. of Gulf and Carib. 14(2):296-302. Ostenfeld, C.H. 1908. On the ecology and distribution of the grass-wrack {Zostera marina) in Danish waters. Report of the Danish Biological Station, Copenhagen, Denmark. Ott, J.A. 1979. Persistence of a seasonal growth rhythm in Posidonia oceanica (L.) Delile under constant conditions of temperature and illumination. Mar. Biol. Lett. 1: 99­ 104. Ozimek, T., E. Pieczynska and A. Hankiewicz. 1991. Effects of filamentous algae on 41: 309-315. submerged macrophyte growth: a laboratory experiment. Aquat. Bot. 1992. Parker, P.L., R.S. Scalan, B.A. Trust and M.C. Woodin. Stable isotopic composition of Redhead Ducks of South Texas and South Dakota. Final Report to U.S. Fish Wildl. Serv. Coop. Agreement No. 14-16-0009-89-917. National Wetlands Research Center. 33 pp. Payne, R.E. 1972. Albedo of the sea surface. J. Atmosph. Sci. 29: 950-967 Pedersen, O. and K. Sand-Jensen. 1993. Water transport in submerged macrophytes. Aquat. Bot. 44: 385-406. P6rez, M. and J. Romero. 1992. Photosynthetic response to light and temperature of the seagrass Cymodoceanodosaandthepredictionofitsseasonality. Aquat.Bot.43:51­ 62. Petersen, CJ.G. 1891. Fiskenes biologiske Forhold i Holbaek Fjord, Rep. Danish Biol. Stn. 1: 1-63. (as in den Hartog 1980). Petersen, CJ.G. 1913. Om Baendeltangens (Zostera marina) Aars-Produktion ide danske Farvande. Mindeskrift for Japetus Steenstrup, pp. 1-20. (as in Rasmussen 1977). 1915. A Petersen, CJ.G. preliminary result ofthe investigation on the valuation ofthe sea. Rep. Danish Biol. Stn, 23: 29-32. (as in Rasmussen 1977). Petersen, CJ.G. 1918. The sea bottom and its production of fish food. A survey of the work done in connection with valuation of the Danish waters from 1883 to 1917. Rep. Danish Biol. Stn. 25: 1-82. Petersen, CJ.G. and P. Boysen-Jensen. 1911. Valuation of the Sea. I. Animal life of the Sea-Bottom, its food and quantitiy. Rep. Danish Biol. Stn. 20: 1-81. Phillips, R.C. 1984. The ecology of eelgrass meadows in the Pacific Northwest: a US Fish Wildl. Serv. 85 community profile. FWS/OBS-84/24. pp. 1978. A mechanism Phillips. G.L., D.Eminson and B. Moss. to account for macrophyte decline in progressively eutrophied freshwaters. Aquat. Bot. 4: 103-126. Phillips, R.C., C. McMillan and K.W. Bridges. 1981. Phenology and reproductive physiology of Thalassia testudinum from the Western Tropical Atlantic. Aquat. Bot. 11: 263-277. Phillips, R.C. and E.G. Menez. 1988. Seagrasses. Smithsonian contributions to Marine Sciences No. 34. Smithsonian Institution Press, Washington, D.C. 104 pp. 1986. Pierce, J.W., D.L. Correll, B. Goldberg, M.A. Faust and W.H. Klein. Response of underwater light transmittance in the Rhode River Estuary to changes in water- quality parameters. Estuaries 9(3): 169-178. Pirc, H. 1985. Growth dynamics in Posidonia oceanica (L.) Delile. I. Seasonal changes of soluble free in carbohydrates, starch amino acids, nitrogen and organic anions , different parts of plant. P.5.Z.N.1.: Mar. Ecol. 6(2): 141-165 1991. above Pohle, D.G., V.M. Bricelj and Z. Garcia-Esquivel. The eelgrass canopy as an bottom refuge from benthic predators for juvenile bay scallops, Argopecten inactions. Mar. Ecol. Prog. Ser. 74: 47-59. Preisendorfer,R.W. 1986. Secchidiskscience:visualopticsofnaturalwaters. Limnol.and Oceanogr. 31(5): 909-926. Pulich, W.M. 1985. Seasonal growth dynamics of Ruppia maritima L. s.l. and Halodule wrightii Aschers. in Southern Texas and evaluation of sediment fertility status. Aquat. Bot. 23: 53-66. 1991. Pulich, W.M. and W.A. White. Decline of submerged vegetation in the Galveston Bay System: chronology and relationships to physical processes. I. Coast. Res. 7(4): 1125-1138. Quammen, M.L. and C.P. Onuf. 1993. Laguna Madre: seagrss changes continue decades after salinity reduction. Estuaries 16(2): 302-310. M.D. Burchett and A. Pulkownik. 1992. Distribution of extractable Ralph, PJ., carbohydrate reserves within the rhizome of the seagrass Posidonia australis Hook, f. Aquat. Bot. 42: 385-392. Rasmussen, E. 1977. The wasting disease of eelgrass (Zostera marina) and its effect on environmental factors and fauna. In: C.P. Mcßoy and C. Helfferich (eds.). Seagrass ecosystems: a scientific perspective. Marcel Dekker, Inc. New York. 1-52 pp. 1980. Raymont, J.E.G. Plankton and productivity in the oceans. Pergamon Press, Oxford. 489 pp. Rice, J.D., R.P. Trocine and G.N. Wells. 1983. Factors influenceing seagrass ecology in the Indian River Lagoon. Florida Sci. 46(3/4): 277-286. Roessler, M.A. and J.C. Zieman. 1969. The effects of thermal additions on the biota in southern Biscayne Bay, Florida. Gulf Carib. Fish. Inst. Proc. 22nd: 136-145. Saffo, M.B. 1987. New light on seaweeds. BioScience 37(9): 654-664 Sand-Jensen, K. 1977. Effects of epiphytes on eelgrass photosynthesis. Aquat. Bot. 3: 55­ 63. 1984. Savastano, K.J., K.H. Faller, and R.L. Iverson. Estimating vegetation coverage in St. airborne Joseph Bay, Florida, with an multispectral scanner. Photogrammetric 50:1159-1170. Engineering and Remote Sensing. ¦ rf Scoffin, T.P. 1970. The trapping and binding of subtidal carbonate sediments by marine vegetation in Bimini Lagoon, Bahamas. J. Sed. Petr. 40(1): 249-273. the distribution Shormann, D.E. 1992. The effects of freshwater inflow and hydrography on of Brown Tide in South Texas Bays. Masters Thesis, University of Texas at Austin. Austin, TX. 11l pp. Short, F.T. 1980. A simulation model of the seagrass production system. In: R.C. Phillips and C.P. Mcßoy (eds.) Handbook of seagrass biology: an ecosystem perspective. Garland STPM Press, New York. 277-295 pp. F.T. 1988. research in the Niantic River. Final to Short, Eelgrass-scallop Report Waterford-East Lyme Shellfish Commission. Unpublished Report. Short, F.T., D.M. Burdick and J.E. Kaldy. submitted. Examining the effects of coastal eutrophication on eelgrass, Zostera marina L., through light reduction and nutrient enrichment in mesocosm systems. Short, F.T., G.E.Jones and D.M. Burdick. 1991. Seagrass decline: problems and solutions. Coastal Zone ’9l Conference-ASCE, Long Beach Ca. July, 1991. Short, F.T. and C.P. Mcßoy. 1984. Nitrogen uptake by leaves and roots of the seagrass Zostera marina L. Bot. Mar. 27: 547-555. Short, F.T., S.W. Nixon and C.A. Oviatt. 1974. Field studies and simulations with a fine grid hydrodynamic model. In: An environmental study of a power plant at nuclear Charlestown, Rhode Island. Univ. of Rhode Island Mar. Tech. Report 33. The - Coastal Resource Center. VIB-1 VIB-27 pp. Short, F.T. and C.A. Short. 1984. The seagrass filter: purification ofestuarine and coastal waters. In: V.S. Kennedy (ed.) The estuary as a filter. Academic Press. 395-413 pp. Simon, J.L. 1974. Tampa Bay estuarine system-a synopsis. Fla. Sci. 37:217-244. Sogard, S.M. and K.W. Able. 1991. A comparison of eelgrass, sea lettuce macroalgae and marsh creeks as habitats for epibenthic fishes and decapods. Est. Coast. Shelf Sci. 33: 501-519. South, G.R. and A. Whittick. 1987. Introduction to phycology. Blackwell Scientific Publications. Oxford London. 341 pp. Staver, K.W. 1985. Responses of epiphytic algae to nitrogen and phosphorus enrichment and effectsonproductivityofthe hostplant,Potamogetonperfoliatus L., inestuarine waters. Masters Thesis. University of Maryland, College Park, MD. 68 pp. Stockwell, D.A., E.J. Buskey and T.E. Whitledge. 1993. Studies on conditions conducive to the development and maintenance of a persistent "Brown tide" in Laguna Madre, Texas. In: TJ. Smayda and Y. Shimizu (eds.) Toxic phytoplankton blooms in the sea. Elsevier Science Publ. 693-698 pp. Taylor, AR.A 1957a. Studies ofthe development ofZostera marina L. I. The embryo and the seed. Can. J. Bot. 35: 477-499 Taylor, A.R.A. 1957b. Studies of the development of Zostera marina L. 11. Germination Can. J. Bot. 35: 681-695. and seedling development. Taylor, G.W., and C.H. Saloman. 1968. Some effects of hydraulic dredging and coastal 67:213-241. development in Boca Ciega Bay, Florida. Fishery Bull. Thayer, G.W., S.M. Adams and M.W. LaCroix. 1975. Structural and functional aspects of a recently established Zostera marina community. In: L.E. Cronin (ed.). Estuarine research. Vol. 1. Academic Press, Inc. New York. 1984. Thayer, G.W., WJ. Kenworthy and M.S. Fonseca. The ecology of eelgrass meadows of the atlantic coast: a community profile. US Fish and Wildl. Serv. FWS/OBS­ -84/02. 147 pp. Thorhaug, A., D. Segar and M.A. Roessler. 1981. Impact of a power plant on a subtropical estuarine environment. Mar. Poll. Bull. 4:166-169. Thresher, R.E., P.D. Nichols, J.S. Gunn, B.D. Bruce and D.M. Furlani. 1992. Seagrass detritus as the basis of a coastal planktonic food chain. limnol. Oceanogr. 37(8): 1754-1758. 1982. Leaf-root interaction in the uptake of ammonium Thursby, G.B. and M.M. Harlin. by Zostera marina. Mar. Biol. 72: 109-112. Tomasko,D.A 1992. Variationingrowthformofshoalgrass(Halodulewrightii)dueto changesinthespectralcomposition oflightbelow acanopyofturtlegrass(Thalassia testudinum). Estuaries 15(2): 214-217. 1991. Tomasko, D.A. and K.H. Dunton. Growth and production of Halodule wrightii in relation to continuous measurementsofunderwater light levels inSouth Texas. In: Kenworthy, WJ. and D.E, Haunert (eds.). The light requirements of seagrasses: of a to examine of water proceedings workshop the capability quality criteria, standards and monitoring programs to protect NOAA Technical seagrasses. Memorandum NMFS-SEFC-287. 79-84 pp. 1985. Nutrient Twilley, R.R., W.M. Kemp, K.W. Staver, J.C. Stevenson and W.R. Boyton. enrichment of estuarine submersed vascular plant communities. I. Algal growth and effects on production of plants and associated communities. Mar. Ecol. Prog. Ser. 23: 179-191. Vincente, V.P. and J.A. Rivera. 1982. Depth limits of the seagrass Thalassia testudinum (Konig) in Jobos and Guayanilla Bays, Puerto Rico. Carib. J. Sci. 17(1-4): 73-79. Virnstein, R.W. and P.A. Carbonara. 1985. Seasonal abundance and distribution of drift in the mid-Indian River Lagoon, Florida. algae and seagrasses Aquat. Bot. 23: 67­ 82. Virnstein, R.W. and M.C. Curran. 1986. Colonization ofartificial seagrass versus time and distance from source. Mar. Ecol. Prog. Ser. 29: 279-288. 1987. Virnstein, R.W. Seagrass-associated invertebrate communities of the Southeastern USA: a reiew. In: M.J. Durako, R.C. Phillips and R.R. Lewis (eds.). Proceedings of the Southeastern United of the symposium on subtropical-tropical seagrasses States. Florida Mar. Res. Publ., Florida Dept. Nat. Res. Number 42. 89-116 pp. Vogt, H. and W. Schramm. 1991. Conspicuous decline of Fucus in Kiel Bay (western 189-194. Baltic): what are the causes? Mar. Ecol. Prog. Ser, 69: 1988. Walker, D.I. and AJ. McComb. Seasonal variation in the production, biomass and nutrient status ofAmphibolis antarctica (Labill.) Sonder ex Aschers. and Posidonia australis Hook. F. in Shark Bay, Western Australia. Aquat. Bot. 31: 259-275. and A.J. McComb. 1989. Effects of boat Walker, D.1., R.J. Lukatelich, G. Bastyan moorings on seagrass beds near Perth, Western Australia. Aquat. Bot. 36:69-77. Ward,G.H. 1979.Hydrographyandcirculationprocessesofgulfestuaries.In:P.Hamilton and K.B. Macdonald (eds). Estuarine and wetland Plenum Press, N.Y. processes. 183-215. pp. Ward, G.H. and N.E. Armstrong. 1980. Matagorda Bay, Texas: its hydrography, ecology and fishery resources. FWS/OBS-81/52. Ward, L.G., M. Kemp and W.R. Boyton. 1984. The influence of waves and seagrass communities on suspended particulates in an estuarine embayment. Mar. Geol. 59: 85-103. Wetzel, R.L. and H.A. Neckles. 1986. A model of Zostera marina L. photosynthesis and growth: simulated effects of selected physical-chemical variables and biological interactions. Aquat. Bot. 26: 307-323. Wetzel, R.L. and P.A. Penhale. 1983. Production ecology of communities in the seagrass lower Chesapeake Bay. Mar. Tech. Soc. J. 17(2): 22-31. 1979. Wiginton, J.R. and C. McMillan. Chlorophyll composition under controlled light conditions as related to the distribution of seagrass in Texas and the U.S. Virgin Islands. Aquat. Bot. 6: 171-184. Williams, S.L. and C.P. Mcßoy. 1982. Seagrass productivity: the effect of light on carbon uptake. Aquat. Bot. 12: 321-344. Williams, S.L. and W.C. Dennison. 1990. light availability and diurnal growth of a green Mar. Biol. macroalga (Caulerpa cupressoides) and a seagrass (Halophila decipiens). 106:437-443. Wood, EJ.F. and J.C. Zieman. 1969. The effects of temperature on estuarine plant communities. Chesapeake Sci. 10(3&4): 172-174. J.C. 1970. The effects of a thermal effluent stress on the and Zieman, seagrasses macroalgae inthevicinityofTurkeyPoint, Biscayne Bay, Florida. Ph.D. Dissertaion, University of Miami, Coral Gables, Florida. Zieman,J.C. 1972.OriginofcircularbedsofThalassia(Spermatophyta:Hydrocharitaceae) in South Biscayne Bay, Florida and their relationship to mangrove hammocks. Bull. Mar. Sci. 22(3): 559-574. Zieman, J.C. 1975a. Tropical seagrass ecosystems and pollution. In: E.J. F. Wood and R.E. Johannes (eds). Tropical marine pollution. Elsevier Scientific Pub. Comp. Amsterdam. 63-74 pp. Zieman, J.C. 1975b. Seasonal variation of turtle grass, Thalassia testudinum Konig, with reference salinity effects. to temperature and Aquat. Bot. 1:107-123. Zieman, J.C. 1976. The ecological effects of physical damage from motor boats on turtle beds in Southern Florida. 127-139. grass Aquat. Bot. 2: 1982. Zieman, J.C. The ecology of the seagrasses of South Florida: a community profile. U.S.Fish andWildl.Serv.,OfficeofBiol.Serv.,WashingtonD.C. FWS/OBS-82/25. 158 pp. Zieman, J.C. and EJ.F. Wood. 1975. Effects of thermal pollution on tropical-type In: Wood estuaries, with emphasis on Biscayne Bay, Florida. EJ.F. and R.E. Johannes marine Elsevier Scientific Pub. (eds.) Tropical pollution. Comp. Amsterdam. 75-98 pp. Zieman, J.C and R.T. Zieman. 1989. The ecology of the meadows of the West seagrass coast of Florida: a community profile. U.S. Fish and Wildl. Serv. Biol. Rep. 85(7.25). 155 pp. R.J. 1976a. The effects of kraft mill effluents Zimmerman, M.S., and Livingston. on benthic macrophyte assemblages in a shallow bay system (Apalachee Bay, north Florida, U.S.A.). Mar. Biol. 34:297-312. 1976b. Zimmerman, M.S., and RJ. Livingston. Seasonality and physico-chemical ranges of benthic macrophytes from a north Florida estuary (Apalachee Bay). Contrib. Mar. Sci. Univ. Texas 20:34-45. Zimmerman, C.F. and J.R. Montgomery. 1984. Effects of a decomposing drift algal mat sediment pore water nutrient concentrations in a Florida bed. Mar. on seagrass Ecol. Prog. Ser. 19: 299-302. Zimmerman, R.C., R.D. Smith and R.S. Alberte. 1987. Is growth of eelgrass nitrogen limited? A numerical simulation of the effects of light and nitrogen on the growth dynamics of Zostera marina. Mar. Ecol. Prog. Ser. 41:167-176. Zimmerman, R.C., J.L. Reguzzoni, S. Wyllie-Echeverria, M. Josselyn and R.S. Alberte. 1991. Assessment of environmental suitability for growth of Zostera marina L. (eelgrass) in San Francisco Bay. Aquat. Bot. 39: 353-366.