diff --git "a/data/datasets/onu/Chapter_06.txt" "b/data/datasets/onu/Chapter_06.txt" new file mode 100644--- /dev/null +++ "b/data/datasets/onu/Chapter_06.txt" @@ -0,0 +1,819 @@ +Chapter 6. Primary Production, Cycling of Nutrients, Surface Layer and Plankton +Writing team: Thomas Malone (Convenor), Maurizio Azzaro, Antonio Bode, Euan Brown Robert Duce, Dan Kamykowski, Sung Ho Kang, Yin Kedong, Michael Thorndyke, an Jinhui Wang, Chul Park (Lead member); Hilconida Calumpong and Peyman Eghtesad (Co-Lead members) +1. Primary Production* +1.1 Definition and ecological significance +Gross primary production (GPP) is the rate at which photosynthetic plants and bacteri use sunlight to convert carbon dioxide (CO2) and water to the high-energy organi carbon compounds used to fuel growth. Free oxygen (O2) is released during the process Net primary production (NPP) is GPP less the respiratory release of CO, b photosynthetic organisms, i.e., the net photosynthetic fixation of inorganic carbon int autotrophic biomass. NPP supports most life on Earth; it fuels global cycles of carbon nitrogen, phosphorus and other nutrients and is an important parameter of atmospheri CO2 and O, levels (and, therefore, of anthropogenic climate change). +Global NPP is estimated to be ~105 Pg C yr’, about half of which is by marine plant (Field et al., 1998; Falkowski and Raven, 1997; Westberry et al., 2008).” Within th euphotic zone of the upper ocean,* phytoplankton and macrophytes‘ respectivel account for ~94 per cent (~50 + 28 Pg C yr’) and ~6 per cent (~3.0 Pg C yr“) of NP (Falkowski et al., 2004; Duarte et al., 2005; Carr et al., 2006; Schneider et al., 2008 Chavez et al., 2011; Ma et al., 2014; Rousseaux and Gregg, 2014). All NPP is not equal i terms of its fate. Marine macrophytes play an important role as carbon sinks in th global carbon cycle, provide habitat for a diversity of animal species, and food fo marine and terrestrial consumers (Smith, 1981; Twilley et al., 1992; Duarte et al., 2005 Duarte et al., 2010; Heck et al., 2008; Nellemann et al., 2009; McLeod et al., 2011 Fourqurean et al., 2012). Phytoplankton NPP fuels the marine food webs upon whic marine fisheries depend (Pauly and Christensen, 1995; Chassot et al., 2010) and the +* Microbenthic, epiphytic and symbiotic algae can be locally important in shallow waters and corals, bu are not addressed here. Chemosynthetic primary production is addressed elsewhere. +21Pg=10¢ +3 Defined here to include the epipelagic (0-200 m) and mesopelagic (200 — -1000 m) zones. The euphoti zone lies within the epipelagic zone. +* Macrophytes include sea grasses, macroalgae, salt marsh plants and mangroves. Phytoplankton ar single -celled, photosynthetic prokaryotic and eukaryotic microorganisms growing in the euphotic zon (the layer between the ocean’s surface and the depth at which photosynthetically active radiation [PAR] i 1 per cent of surface intensity). Most phytoplankton species are > 1 um and < 1 mm in equivalen spherical diameter (cf. Ward et al., 2012). +© 2016 United Nations + +“biological pump” which transports 2-12 Pg C yr* of organic carbon to the deep se (Falkowski et al., 1998; Muller-Karger et al., 2005; Emerson and Hedges, 2008; Doney 2010; Passow and Carlson, 2012), where it is sequestered from the atmospheric pool o carbon for 200-1500 years (Craig, 1957; Schlitzer et al., 2003; Primeau and Holzer, 2006 Buesseler, et al., 2007). +Changes in the size structure of phytoplankton communities influence the fate of NP (Malone, 1980; Legendre and Rassoulzadegan, 1996; Pomeroy et al., 2007; Marafion 2009). In general, small cells (picophytoplankton with equivalent spherical diameters < um) account for most NPP in subtropical, oligotrophic (< 0.3 mg chlorophyll-a m*) nutrient-poor (nitrate + nitrite < 1 uM), warm (> 20°C) waters. Under these conditions the flow of organic carbon to harvestable fisheries and the biological pump are relativel small. In contrast, larger cells (microphytoplankton > 20 um) account for > 90 per cent o NPP in more eutrophic (> 5 mg chlorophyll-a m°), nutrient-rich (nitrate + nitrite >1 uM), cold (< 15°C) waters (Kamykowski, 1987; Agawin et al., 2000; Buitenhuis et al. 2012). Under these conditions, diatoms® account for most NPP during spring blooms a high latitudes and periods of coastal upwelling when NPP is high and nutrients are no limiting (Malone, 1980). The flow of organic carbon to fisheries and the biological pum is higher when larger cells account for most NPP (Laws et al., 2000; Finkel et al., 2010). +1.2 Methods of measuring net primary production (NPP 1.2.1 Phytoplankton Net Primary Production +Phytoplankton (NPP) has been estimated using a variety of in situ and remote sensin methods (Platt and Sathyendranath, 1993; Geider et al., 2001; Marra, 2002; Carr et al. 2006; Vernet and Smith, 2007; Cullen, 2008a; Cloern et al., 2013). Multiplatform (e.g. ships, moorings, drifters, gliders, aircraft, and satellites) sampling strategies that utiliz both approaches are needed to effectively detect changes in NPP on ecosystem t global scales (UNESCO-IOC, 2012). +On small spatial and temporal scales (meters-kilometres, hours-days), severa techniques have been used including oxygen production and the incorporation of "° and “Cc labelled bicarbonate (Cullen, 2008a). The most widely used and standar method against which other methods are compared or calibrated is based on th incorporation of ““C-bicarbonate into phytoplankton biomass (Steeman-Nielsen, 1963 Marra, 1995; Marra, 2002; Vernet and Smith, 2007; Cullen, 2008a). On large spatia scales (Large Marine Ecosystems’ to the global ocean), the most effective way to detec space-time variability is via satellite-based measurements of water-leaving radianc combined with diagnostic models of depth-integrated NPP as a function of depth- +° Diatom growth accounts for roughly half of marine NPP and therefore for about a quarter of globa photosynthetic production (Smetacek, 1999). +8 Large marine ecosystems (200,000 km? or larger) are coastal ecosystems characterized by their distinc bathymetry, hydrography, productivity and food webs (Sherman et al., 1993). +© 2016 United Nations + +integrated chlorophyll-a concentration (W Chl), photosynthetically active solar radiation and temperature (Antoine and Morel, 1996; Perry, 1986; Morel and Berthon, 1989; Plat and Sathyendranath, 1993; Behrenfeld and Falkowski, 1997; Sathyendranath, 2000 Gregg et al., 2003; Behrenfeld et al., 2006; Carr et al., 2006; Arrigo et al., 2008; Bissinge et al., 2008; McClain, 2009; Westberry et al., 2008; Cullen et al., 2012; Siegel et al. 2013). +An overview of the latest satellite based models may be found at the Ocean Productivit website.’ Satellite ocean-colour radiometry (OCR) data have been used to estimate surfac chlorophyll-a fields and NPP since the Coastal Zone Color Scanner (CZCS) mission (1978 1986). Uninterrupted OCR measurements began with the Sea-viewing Wide Field-of-vie Sensor (SeaWiFS) mission (1997-2010) (Hu et al., 2012). A full accounting of current pola orbiting and geostationary ocean-colour sensors with their capabilities (swath width, spatia resolution, spectral coverage) can be found on the web site of the International Ocean Colour Coordinating Group.® +The skill of model-based estimates of NPP has been improving (O'Reilly et al., 1998; Lee 2006; Friedrichs et al., 2009; Saba et al., 2010; Saba et al., 2011; Mustapha et al., 2012) but further improvements are needed through more accurate estimates of W Chl Chlorophyll-a fields can be estimated more accurately by blending data from remot sensing and in situ measurements, especially in regions where in situ measurements ar sparse and in turbid, coastal ecosystems (Conkright and Gregg, 2003; Gregg et al., 2003 Onabid, 2011). An empirical approach has been developed for ocean-colour remot sensing called Empirical Satellite Radiance-In situ Data (ESRID) algorithm (Gregg et al. 2009). +1.2.2. Macrophyte Net Primary Production +The NPP of macroalgae, sea grasses, salt marsh plants and mangroves can be estimate by sequentially (e.g., monthly during the growing season) measuring increases i biomass (including leaf litter in salt marshes and mangrove forests) using a combinatio of in situ techniques (e.g., Mann, 1972; Cousens, 1984; Dame and Kenny, 1986 Amarasinghe and Balasubramaniam, 1992; Long et al., 1992; Day et al., 1996; Ross et al. 2001; Curco et al., 2002; Morris, 2007) and satellite-based multispectral imagery (e.g. Gross et al., 1990; Zhang et al., 1997; Kovacs et al., 2001; Gitelson, 2004; Liu et al., 2008 Kovacs et al., 2009; Heumann, 2011; Mishra et al., 2012; Son and Chen, 2013). Fo remote sensing, accurate in situ measurements are critical for validating models used t map these habitats and estimate NPP (Gross et al., 1990; Kovacs et al., 2009; Roelfsem et al., 2009; Mishra et al., 2012; Jia et al., 2013; Trilla et al., 2013). These include shoot or leaf-tagging techniques, measurements of “C incorporation into leaves, an measurements of dissolved O2 production during the growing season (Bittaker and +” http://www.science.oregonstate.edu/ocean.productivity/ 8 http://www.ioccg.org/sensors/current.html. +© 2016 United Nations + +Iverson, 1976; Kemp et al., 1986; Duarte, 1989; Kaldy and Dunton, 2000; Duarte an Kirkman, 2001; Plus et al., 2001, Silva et al., 2009). +1.2.3 The Phenology’ of Phytoplankton Annual Cycles +The timing of seasonal increases in phytoplankton NPP is determined by environmenta parameters, including day length, temperature, changes in vertical stratification, and th timing of seasonal sea-ice retreat in polar waters. All but day length are affected b climate change. Thus, phytoplankton phenology provides an important tool fo detecting climate-driven decadal variability and secular trends. Phenological metrics t be monitored are the time of bloom initiation, bloom duration and time of maximu amplitude (Siegel et al., 2002; Platt et al., 2009). +1.3 Spatial patterns and temporal trends +Marine NPP varies over a broad spectrum of time scales from tidal, daily and seasona cycles to low-frequency basin-scale oscillations and multi-decade secular trend (Malone, 1971; Pingree et al., 1975; Steele, 1985; Cloern, 1987; Cloern, 2001; Cloern e al., 2013; Duarte, 1989; Powell, 1989; Malone et al., 1996; Henson and Thomas, 2007 Vantrepotte and Mélin, 2009; Cloern and Jassby, 2010; Bode et al., 2011; Chavez et al. 2011). Our focus here is on low-frequency cycles and multi-decade trends. +1.3.1 Phytoplankton NPP +For the most part, the global pattern of phytoplankton NPP (Figure 1) reflects th pattern of deep-water nutrient inputs to the euphotic zone associated with winte mixing and thermocline erosion at higher latitudes, thermocline shoaling at lowe latitudes, and upwelling along the eastern boundaries of the ocean basins and th equator (Wollast, 1998; Pennington et al., 2006; Chavez et al., 2011; Ward et al., 2012) The global distribution of phytoplankton NPP is also influenced by iron limitation an grazing by microzooplankton in so-called High Nutrient Low Chlorophyll (HNLC) zone which account for ~20 per cent of the global ocean, e.g., oceanic waters of the subarcti north Pacific, subtropical equatorial Pacific, and Southern Ocean (Martin et al., 1994 Landry et al., 1997; Edwards et al., 2004). Nutrient inputs associated with river runof enhance NPP in coastal waters during the growing season (Seitzinger et al., 2005 Seitzinger et al., 2010). Annual cycles of NPP associated with patterns of nutrient suppl and seasonal variations in sunlight tend to increase in amplitude and decrease i duration with increasing latitude. Seasonal increases in NPP generally follow winte mixing when nutrient concentrations are high, the seasonal thermocline sets up, an day length increases. Annual cycles are also more pronounced in coastal waters subjec to seasonal upwelling. +° Phenology is the study of the timing and duration of cyclic and seasonal natural phenomena (e.g., sprin phytoplankton blooms, seasonal cycles of zooplankton reproduction), especially in relation to climate an plant and animal life cycles. +© 2016 United Nations + +The boundaries and names shown and the designations used on this map do not imply official endorsement or acceptance by the United Nations. +Figure 1. Climatological map Distribution of annual marine NPP for (a) NASA Ocean Biogeochemical Mode and (b) Vertically-Integrated Production Model (VGPM) for the period from September 1998 to 201 (Rousseaux — August 1999 (Blue < 100 gC m7, Green > 110 gc m” and < 400 gc m?, Red > 400 gc m” (Rutgers Institute of Marine and Gregg, 2014). Globally, diatoms accounted for about 50 per cent of NP while coccolithophores, chlorophytes and cyanobacteria accounted for about 20 per cent, 20 per cent an 10 per cent, respectively. Diatom NPP was highest at high latitudes and in equatorial and easter boundary upwelling systems. Coastal Sciences, http://marine.rutgers.edu/opp/). Coastal ecosystems (re — green) and the permanently stratified subtropical waters of the central gyres (blue) each account fo ~30 per cent of the ocean’s NPP, whereas the former accounts for only ~8 per cent of the ocean’s surfac area compared to ~60 per cent for the open ocean waters of the subtropics (Geider et al., 2001; Marafid et al., 2003; Muller-Karger et al., 2005). +© 2016 United Nations + +The global distribution of annual NPP in the ocean can be partitioned into broa provinces with eastern boundary upwelling systems and estuaries exhibiting the highes rates and subtropical central gyres the lowest rates (Figure 1, Table 1). +Table 1. Ranges of phytoplankton mean daily NPP and annual NPP reported for different marin provinces. Estimates are based on in situ measurements and models using satellite-based observations o chlorophyll-a fields. Western boundaries of the ocean basins generally feature broad continental shelve and eastern boundaries tend to have narrow shelves with coastal upwelling. (Data sources: Malone et al. 1983; O’Reilly and Busch, 1983; Pennock and Sharp, 1986; Cloern, 1987; Malone, 1991; Barber et al. 1996; Karl et al., 1996; Malone et al., 1996; Pilskaln et 173 al., 1996; Smith and DeMaster, 1996; Lohren et al., 1997; Cloern, 2001; Smith et al., 2001; Steinberg et al., 2001; Marafion et al., 2003; Sakshaug, 2004 PICES, 2004; Teira et al., 2005; Tian et al., 2005; Pennington et al., 2006; Subramanian et al., 2008; Verne et al., 2008; Bidigare et al., 2009; Sherman and Hempel, 2009; Chavez et al., 2010; 176 Saba et al., 2011 Brown and Arrigo, 2012; Cloern et al., 2013; Lomas et al., 2013). +Province mg C md" gc m” y Subtropical Central Gyres 20 - 1,040 150-17 Western Boundaries 10 - 3,500 200 - 47 Eastern Boundaries 30 - 7,300 460 -— 1,25 Equatorial Upwelling 640 - 900 24 Arctic Ocean 3-1,100 5-40 Southern Ocean 290 — 370 50-45 Coastal Seas 100 — 1,400 40 - 60 Estuaries & Coastal Plumes 100 — 8,000 70 - 1,890 +Interannual variability and multi-decadal trends in phytoplankton NPP on regional t global scales are primarily driven by: (1) climate change (e.g., basin-scale oscillations an decadal trends, including loss of polar ice cover, upper ocean warming, and changes i the hydrological cycle); (2) land-based, anthropogenic nutrient loading; and (3) pelagi and benthic primary consumers. Global-scale trends in phytoplankton NPP remai controversial (Boyce et al., 2010; Boyce et al., 2014; Mackas, 2011; Rykaczewski an Dunne, 2011; McQuatters-Gollop et al., 2011; Dave and Lozier, 2013; Wernand et al. 2013).). Remote sensing (sea-surface temperature and chlorophyll fields), mode simulations and marine sediment records suggest that global phytoplankton NPP ma have increased over the last century as a consequence of basin-scale climate forcin that promotes episodic and seasonal nutrient enrichment of the euphotic zone throug vertical mixing and upwelling (McGregor et al., 2007; Bidigare et al., 2009; Chavez et al. 2011; Zhai et al., 2013). In contrast, global analyses of changes in chlorophyl distribution over time suggest that annual NPP in the global ocean has declined over the +© 2016 United Nations + +last 100 years (Gregg et al., 2003; Boyce et al., 2014). A decadal scale decline i consistent with model simulations indicating that both NPP and the biological pum have decreased by ~7 per cent and 8 per cent, respectively, over the last five decade (Laufkétter et al., 2013), trends that are likely to continue through the end of thi century (Steinacher et al., 2010). +Given uncertainties concerning global trends, long-term impacts of secular changes i phytoplankton NPP on food security and climate change cannot be assessed at this tim with any certainty. Resolving this controversy and predicting future trends will requir sustained, multi-decadal observations and modelling of phytoplankton NPP and ke environmental parameters (e.g., upper ocean temperature, pCO, pH, depth of th aragonite saturation horizon, vertical stratification and nutrient concentrations) o regional and global scales — observations that may have to be sustained for at leas another 40-50 years (Henson et al., 2010). +1.3.2 Macrophyte NPP +Marine macrophyte NPP, which is limited to tidal and relatively shallow waters i coastal ecosystems, varies from 30-1,200 g C m® yr“ (Smith, 1981; Charpy-Roubaou and Sournia, 1990; Geider et al., 2001; Duarte et al., 2005; Duarte et al., 2010 Fourqurean et al., 2012; Ducklow et al., 2013). In contrast to the uncertainty of decada trends in phytoplankton NPP, decadal declines in the spatial extent and biomass o macrophytes (a proxy for NPP) over the last 50-100 years are relatively wel documented. Macrophyte habitats are being lost and modified (e.g., fragmented) a alarming rates (Duke et al., 2007; Valiela et al., 2009; Waycott et al., 2009; Wernberg e al., 2011), i.e., 2 per cent for macrophytes as a group, with total areal losses to date o 29 per cent for seagrasses, 50 per cent for salt marshes and 35 per cent for mangrov forests (Valiela et al., 2001; Hassan et al., 2005; Orth et al., 2006; Waycott et al., 2009 Fourqurean et al., 2012). As a whole, the world is losing its macrophyte ecosystems i coastal waters four times faster than its rain forests (Duarte et al., 2008), and the rate o loss is accelerating (Waycott et al., 2009). +2. Nutrient Cycles +Nitrogen (N) and phosphorus (P) are major nutrients required for the growth of al organisms, and NPP is the primary engine that drives the cycles of N and P in the oceans The cycles of C, N, P and O, are coupled in the marine environment (Gruber, 2008). A discussed in section 6.1.3, the global pattern of phytoplankton NPP reflects the patter of dissolved inorganic N and P inputs to the euphotic zone from the deep ocean (Figur 1). Superimposed on this pattern are nutrient inputs associated with N fixation atmospheric deposition, river discharge and submarine ground water discharge. I regard to the latter, ground water discharge may be a significant source of N locally i some parts of Southeast Asia, North and Central America, and Europe, but on the scal of ocean basins and the global ocean, ground water discharge of N has been estimated +© 2016 United Nations + +to be on the order of 2-4 per cent of river discharge (Beusen et al., 2013). Given this and challenges of quantifying ground water inputs on ocean basin to global scales (NRC 2004), this source is not considered herein. +2.1 Nitrogen +The ocean's nitrogen cycle is driven by complex microbial transformations, including fixation, assimilation, nitrification, anammox and denitrification (Voss et al., 2013 (Figure 2). NPP depends on the supply of reactive N (N,)’°to the euphotic zone Although most dissolved chemical forms of N, can be assimilated by primary producers the most abundant chemical form, dissolved dinitrogen gas (Nz), can only be assimilate by marine diazotrophs.*" N, inputs to the euphotic zone occur via fluxes of nitrate fro deep water (vertical mixing and upwelling), marine No fixation, river discharge, an atmospheric deposition.’2 N, is removed from the marine N inventory throug denitrification and anammox”™ with subsequent efflux of N. and N20 to the atmospher (Thamdrup et al., 2006; Capone, 2008; Naqvi et al., 2008; Ward et al., 2009; Ward, 2013) Although there is no agreement concerning the oxygen threshold that defines th geographic extent of denitrification and anammox (Paulmier and Ruiz-Pino, 2009), thes processes are limited to suboxic waters with very low oxygen concentrations (< 22 uM). +© Reactive or fixed N forms include dissolved inorganic nitrate, nitrite, ammonium and organic compounds, such as urea and free amino acids. +™ Prokaryotic, free -living and symbiotic bacteria, cyanobacteria and archaea. +” River discharge and atmospheric deposition include nitrate from fossil fuel burning and fixed N i synthetic fertilizer produced by the Haber-Bosch process for industrial nitrogen fixation. +*8 Anaerobic ammonium oxidation. +© 2016 United Nations + +Nz Atmosphere +Nitrification +PON ——~NH,* —> NO, — > NO, +Suboxic NO; +| +Organic NO, +Nitrogen —_1__ +Remineralization +Figure 2. The biological nitrogen cycle showing the main inorganic forms in which nitrogen occurs in th ocean (PON-pariculate organic nitrogen) (adapted from Ward, 2012). +Variations in the ocean’s inventory of N, have driven changes in marine NPP an atmospheric CO, throughout the Earth’s geological history (Falkowski, 1997; Gruber 2004; Arrigo, 2005). Marine N> fixation provides a source of “new” N and NPP that fue marine food webs and the biological pump. Thus, the rate of N> fixation can affec atmospheric levels of CO2 on time-scales of decades (variability in upper ocean nutrien cycles) to millennia (changes in the N, inventory of the deep sea). This makes th balance between the conversion of N2 to biomass (Nz fixation) and the production of N (reduction of nitrate and nitrite by denitrification and anammox) particularly importan processes in the N cycle that govern the marine inventory of N, and sustain life in th oceans (Karl et al., 2002; Ward et al., 2007; Gruber, 2008; Ward, 2012). +2.1.1 The Marine Nitrogen Budget +Estimates of global sources and sinks of N, vary widely (Table 2). Marine biological N fixation accounts for ~50 per cent of Nz fixation globally (Ward, 2012). Most marine N fixation occurs in the euphotic zone of warm (> 20°C), oligotrophic waters between 30 N and 30° S (Karl et al., 2002; Mahaffey et al., 2005; Stal, 2009; Sohm et al., 2011) Denitrification and anammox in benthic sediments and mid-water oxygen minimum +© 2016 United Nations + +zones (OMZs) account for most losses of N from the marine N, inventory (Ulloa et al. 2012; Ward, 2013). +Table 2. Summary of estimated sources and sinks (Tg N yr’) in the global marine nitrogen budget. (Dat sources: Codispoti et al., 2001; Gruber and Sarmiento, 2002; Karl et al., 2002; Galloway et al., 2004 Mahaffey et al., 2005; Seitzinger et al., 2005; Boyer et al., 2006; Moore et al., 2006; Deutsch et al., 2007 Duce et al., 2008; DeVries et al., 2012; Grosskopf et al., 2012; Luo et al., 2012; Naqvi, 2012.) +Sources | N fixation 60-20 Rivers 35-8 Atmosphere 38-9 TOTAL 133-376 +Sinks Denitrification & anammox | 120-450 +Sedimentation 2 N20 loss 4- TOTAL 149-482 +Assuming a C:N:P ratio of 106:16:1 (the Redfield Ratio, Redfield et al., 1963), th quantity of N, needed to support NPP globally is ~8800 Tg N yr’. Given curren estimates, inputs of N, from river discharge and atmospheric deposition support 2-4 pe cent of NPP annually, i.e., most NPP is supported by recycled nitrate from deep water (cf. Okin et al., 2011). +Although the N20 flux is a small term in the marine N budget (Table 2), it is a significan input to the global atmospheric N20 pool. Given a total input of 17.7 Tg N yr- (Freing e al., 2012), marine sources may account for 20-40 per cent of N2O inputs to th atmosphere. As N20 is 200-300 times more effective than CO as a greenhouse gas increases in N,O from the ocean may contribute to both global warming and th destruction of stratospheric ozone. We note that although global estimates fo anammox have yet to be made, this anaerobic process may be responsible for most N production in some oxygen minimum zones (OMZs) (Strous et al., 2006; Hamersley et al. 2007; Lama et al., 2009; Koeve and Kahler, 2010; Ulloa et al., 2012). +The accounting in Table 2 suggests that total sinks may exceed total sources, but th difference is not significant. Many scientists believe that biological N2 fixation i underestimated or the combined rates of denitrification and anammox ar overestimated (Capone, 2008). On average, the Redfield Ratio approximates the C:N: ratio of phytoplankton biomass, and the distribution of deviations from the Redfiel Ratio (Martiny et al., 2013) suggests that: sources exceed sinks in the subtropical gyres sources and sinks are roughly equal in upwelling systems (including their OMZs); an sources tend to be less than sinks at high latitudes. This pattern is consistent with the +© 2016 United Nations 1 + +known distribution of marine diazotrophs and the observation that most marine N fixation occurs in warm, oligotrophic waters between 30° N and 30° S (Mahaffey et al. 2005; Stal, 2009; Sohm et al., 2011). However, given the wide and overlapping ranges o current estimates of N, sources and sinks (Table 2), the extent to which the two are i steady state remains controversial. +Atmospheric deposition of iron to the oceans via airborne dust may ultimately contro the rate of N> fixation in the global ocean and may account for the relatively high rate o N>2 fixation in the subtropical central gyres (Karl et al., 2002). Fe Il is required fo photosynthetic and respiratory electron transport, nitrate and nitrite reduction, and N fixation. The large dust plume that extends from North Africa over the subtropical Nort Atlantic Ocean dominates the global dust field (Stier et al., 2005). Consequently, iro deposition is particularly high in this region (Mahowald et al., 2005) where it ma increase phytoplankton NPP by stimulating N» fixation (Mahowald et al., 2005 Krishnamurthy et al., 2009; Okin et al., 2011). Model simulations indicate that th distribution and rate of N2 fixation may also be influenced by non-Redfield uptake of and P by non-N; fixing phytoplankton (Mills and Arrigo, 2010). In these simulations, N fixation in ecosystems dominated by phytoplankton with N:P ratios < Redfield is lowe than expected when estimated rates are based on Redfield stoicheiometry. In contrast in systems dominated by phytoplankton with N:P ratios > Redfield, N2 fixation is highe than expected based on Redfield stoicheiometry. +2.1.2 Time-Space Coupling of N2 Fixation and Denitrification/Anammox +Early measurements of N2 fixation and the geographic distribution of in situ deviation from the Redfield Ratio suggest that the dominant sites of No fixation and denitrificatio are geographically separated and coupled on the time scales of ocean circulatio (Capone et al., 2008 and references therein). In this scenario, the ocean oscillate between being a net source and a net sink of N, on time scales of hundreds to thousand of years (Naqvi, 2012). However, there is also evidence that N, fixation is closely couple with denitrification/anammox in upwelling-OMZ systems”, i.e., rates of No fixation ar high downstream from OMZs where denitrification/anammox is high (Deutsch et al. 2007). Their findings indicate that N2 fixation and denitrification are in steady state on global scale. Results from 3-D inverse modelling (DeVries et al., 2013) and observation that the marine N, inventory has been relatively stable over the last several thousan years (Gruber, 2004; Altabet, 2007) support the hypothesis that rates of N2 fixation an denitrification/anammox are closely coupled in time and space. +At the same time, global biogeochemical modelling suggests that the negativ feedbacks stabilizing the N, inventory cannot persist in an ocean where N> fixation an denitrification/anammox are closely coupled, i.e., spatial separation, rather than spatia proximity, promoted negative feedbacks that stabilized the marine N inventory and +* Oxygen minimum zones (OMZs) are oxygen-deficient layers in the ocean's water column (Paulmier an Ruiz-Pino, 2009). +© 2016 United Nations 1 + +sustained a balanced N budget (Landolfi et al., 2013). If the coupling is close as argue above, the budget may not be in steady state. In this scenario, increases in vertica stratification of the upper ocean and expansion of OMZs associated with ocean warmin (Keeling et al., 2010) could lead to closer spatial coupling of N» fixation an denitrification, a net loss of N from the marine N, inventory, and declines in NPP an CO2 sequestration during this century. +2.2 Phosphorus +Phosphorus (P) is an essential nutrient utilized by all organisms for energy transport an growth. The primary inputs of P occur via river discharge and atmospheric depositio (Table 3). Biologically active P (BAP) in natural waters usually occurs as phosphate (PO 3), which may be in dissolved inorganic forms (including orthophosphates an polyphosphates) or organic forms (organically bound phosphates). Natural inputs of BA begin with chemical weathering of rocks followed by complex biogeochemica interactions, whose time scales are much longer than anthropogenic P inputs (Benitez Nelson, 2000). Primary anthropogenic sources of BAP are industrial fertilizer, sewag and animal wastes. +The Marine Phosphorus Budget: River discharge of P into the coastal ocean accounts fo most P input to the ocean (Table 3). However, most riverine P is sequestered i continental shelf sediments (Paytan and McLaughlin, 2007) so that only ~25 per cent o the riverine input enters the NPP-driven marine P cycle. Estimates of BAP reaching th open ocean from rivers range from a few tenths to perhaps 1 Tg P yr’ (Seitzinger et al. 2005; Meybeck, 1982; Sharpies et al., 2013). Mahowald et al., (2008) estimated tha atmospheric inputs of BAP are ~0.1 Tg P yr~. Together these inputs would support ~0. per cent of NPP annually. Thus, like Nr, virtually all NPP is supported by BAP recycle within the ocean on a global scale. +Table 3. Summary of estimated sources and sinks (Tg P yr-1) in the global marine phosphorus budget (Data sources: Filippelli and Delaney, 1996; Howarth et al., 1996; Benitez-Nelson, 2000; Compton et al., +2000; Ruttenberg, 2004; Seitzinger et al., 2005; Paytan and McLaughlin, 2007; Mahowald et al., 2008 Harrison et al., 2010; Krishnamurthy et al., 2010.) +Sources | River discharge 10.79 — 31.0 Atmospheric deposition 0.54-1.0 TOTAL 11.33 — 32.05 +Sinks Open ocean sedimentation | 1.30 — 10.57 +The primary source of P in the atmosphere is mineral dust, accounting for approximatel 80 per cent of atmospheric P. Other important sources include biogenic particles biomass burning, fossil-fuel combustion, and biofuels. The P in mineral particles is not +© 2016 United Nations 1 + +very soluble, and most of it is found downwind of desert and arid regions. Only ~0.1 Tg yr? of BAP appears to enter the oceans via atmospheric deposition (Mahowald et al. 2008). Although a small term in the P budget (Table 3), atmospheric deposition appear to be the main external source of BAP in the oligotrophic waters of the subtropical gyre and the Mediterranean Sea (Paytan and McLaughlin, 2007; Krishnamurthy et al., 2010). +Burial in continental shelf and deep-sea sediments is the primary sink, with mos riverine input being removed from the marine P cycle by rapid sedimentation o particulate inorganic (non-reactive mineral lattices) P in coastal waters (Paytan an McLaughlin, 2007). Burial in deep-sea sediments occurs after transformations fro dissolved to particulate forms in the water column. Of the riverine input, 60-85 per cen is buried in continental shelf sediments (Slomp, 2011). Assuming that inputs from rive discharge and atmospheric deposition are, respectively, ~15 Tg P yr’ and 1 Tg P yr’, an that 11 Tg P yr’ and 5 Tg P yr’, respectively, are buried in shelf and open-ocea sediments, the P budget appears to be roughly balanced on the scale of P turnove times in the ocean (~1500 years, Paytan and McLaughlin, 2007). +3. Variability and Resilience of Marine Ecosystems +3.1 Phytoplankton species diversity and resilience +Biodiversity enhances resilience by increasing the range of possible responses t perturbations and the likelihood that species will functionally compensate for on another following disturbance (functional redundancy) (McCann, 2000; Walker et al. 2004; Hooper et al., 2005; Haddad et al., 2011; Appeltans et al., 2012; Cleland, 2011) Annually averaged phytoplankton species diversity of the upper ocean tends to b lowest in polar and subpolar waters, where fast-growing (opportunistic) species accoun for most NPP, and highest in tropical and subtropical waters, where small phytoplankto (< 10 um) account for most NPP (Barton et al., 2010). Phytoplankton species diversity i also a unimodal function of phytoplankton NPP, with maximum diversity at intermediat levels of NPP and minimum diversity associated with blooms of diatoms, dinoflagellates Phaeocystis sp., and coccolithophores (Irigoien et al., 2004). This suggests that pelagi marine food webs may be most resilient to climate and anthropogenic forcings a intermediate levels of annual phytoplankton NPP. +3.2 Events, phenomena and processes of special interest +Zooplankton grazing: Zooplankton populations play key roles in both microbial foo webs” supported by small phytoplankton (< 10 um) and metazoan food webs’® +* The microbial food web (or microbial loop) consists of small phytoplankton (mean spherical diameter 10 um), heterotrophic bacteria, archaea and protozoa (flagellates and ciliates). +© 2016 United Nations 1 + +supported by large phytoplankton (> 20 um). As such, they are critical links in nutrien cycles and the transfer of NPP to higher trophic levels of metazoan consumers. They fue the biological pump and they limit excessive increases in NPP (e.g., Corten and Linley 2003; Greene and Pershing, 2004; Steinberg et al., 2012). Microbial food webs dominat the biological cycles of C, N and P in the upper ocean and feed into metazoan food web involving zooplankton, planktivorous fish, and their predators (Pomeroy et al., 2007 Moloney et al., 2011; Ward et al., 2012). Zooplankton in microbial food webs ar typically dominated by heterotrophic and mixotrophic flagellates and ciliates. Metazoa food webs dominate the flow of energy and nutrients to harvestable fish stocks and t the deep sea (carbon sequestration). Zooplankton in metazoan food webs are typicall dominated by crustaceans (e.g., copepods, krill and shrimp) and are part of relativel short, efficient, and nutritionally rich food webs supporting large numbers o planktivorous and piscivorous fish, seabirds, and marine mammals (Richardson, 2008 Barnes et al., 2010; Barnes et al., 2011). +Microbial food webs support less zooplankton biomass than do metazoan food webs and a recent analysis suggests that zooplankton/phytoplankton ratios range from a lo of ~0.1 in the oligotrophic subtropical gyres to a high of ~10 in upwelling systems an subpolar regions (Ward et al., 2012). Such a gradient is consistent with a shift fro “bottom-up”, nutrient-limited NPP in the oligotrophic gyres, where microflagellates ar the primary consumers of NPP (Calbet, 2008), to “top-down”, grazing control of NPP b zooplankton in more productive high-latitude and upwelling ecosystems, wher planktonic crustaceans are the primary grazers of NPP (Ward et al., 2012). Thus zooplankton grazing on phytoplankton is an important parameter of spatial patterns an temporal trends in NPP, particularly at high latitudes and in coastal upwelling system (section 6.1.4). +3.2.1 NPP and Fisheries +Fish production depends to a large extent on NPP but the relationship between NPP an fish landings is complex. For instance, Large Marine Ecosystems (LMEs) of the coasta ocean account for ~30 per cent of marine phytoplankton NPP and ~80 per cent o marine fish landings globally (Sherman and Hempel, 2009). They are also “provin grounds” for the development of ecosystem-based approaches (EBAs) to fisherie management (McLeod and Leslie, 2009; Sherman and Hempel, 2009; Malone et al. 2014b). EBAs are guided in part by the recognition that the flow of energy and nutrient from NPP through marine food webs ultimately limits annual fish landings (Pauly an Christensen, 1995; Pikitch et al., 2004). +Both mean annual and maximum fish landings have been shown to be related to NPP o regional scales, with increases in potential landings at high latitudes (> 30 per cent) an decreases at low latitudes (up to 40 per cent) (Pauly and Christensen, 1995; Ware, 2000; +*© The so-called “classical” food web is dominated by larger phytoplankton, metazoan zooplankton an nekton. +© 2016 United Nations 1 + +Ware and Thomson, 2005; Frank et al., 2006; Chassot et al., 2007; Sherman and Hempel 2009; Blanchard et al., 2012). However, the NPP required to support annual fis landings (PPR) varies among LMEs, e.g., fisheries relying on NPP at the Eastern Boundar Upwelling Systems require substantially higher levels of NPP than elsewhere (Chassot e al., 2010). Variations in PPR/NPP are related to a number of factors, including th relative importance of microbial and metazoan food webs and differences in th efficiencies of growth and transfer efficiencies among trophic levels. The level o exploitation (PPR/NPP) increased by over 10 per cent from 2000 to 2004, and the NP appropriated by current global fisheries is 17-112 per cent higher than tha appropriated by sustainable fisheries. Temporal and spatial variations in PPR/NPP cal into question the usefulness of global NPP per se as a predictor of global fish landing (Friedland et al., 2012). Friedland et al. (2012) found that NPP is a poor predictor of fis landings across 52 LMEs, with most variability in fish landings across LMEs accounted fo by chlorophyll-a concentration, the fraction of NPP exported to deep water, and th ratio of secondary production to NPP. Given these considerations and uncertaintie concerning the effects of climate change on fluxes of nutrients to the euphotic zone, it i not surprising that there is considerable uncertainty associated with projections of ho changes in NPP will affect fish landings over the next few decades. +3.2.2 NPP Fisheries and zooplankton +Zooplankton is a critical link between NPP and fish production (Cushing, 1990 Richardson, 2008). Efficient transfer of phytoplankton NPP to higher trophic level ultimately depends on the relative importance of microbial and metazoan foods web and the coherence between the timing of phytoplankton blooms (initiation, amplitude duration) and the reproductive cycles of zooplankton and planktivorous fish (Cushing 1990; Platt et al., 2003; Koeller et al., 2009; Jansen et al., 2012). Energy transfer t higher trophic levels via microbial food webs is less efficient than for metazoan foo webs (e.g., Barnes et al., 2010; Barnes et al., 2011; Suikkanen et al., 2013). Coherence i time and space is especially important in higher-latitude ecosystems (Sherman et al. 1984; Edwards and Richardson, 2004; Richardson, 2008; Ohashia et al., 2013), wher seasonal variations in NPP are most pronounced and successful fish recruitment is mos dependent on synchronized production across trophic levels (Cushing, 1990; Beaugran et al., 2003). The phenological response to ocean warming differs among functiona groups of plankton, resulting in predator-prey mismatches that may influence PPR/NP in marine ecosystems. For example, phytoplankton blooms in the North Atlantic begi earlier south of 40°N (autumn — winter) and in spring north of 40°N (Siegel et al., 2002 Ueyama and Monger, 2005; Vargas et al., 2009). Likewise, a 44-year time series (1958 2002) revealed progressively earlier peaks in abundance of dinoflagellates (23 days) diatoms (22 days) and copepods (10 days) under stratified summer conditions in th North Sea (Edwards and Richardson, 2004). Such differential responses in phytoplankto and zooplankton phenology lead to mismatches between successive trophic levels and therefore, a decline in PPR/NPP, i.e., a decrease in carrying capacity for harvestable fis stocks. +© 2016 United Nations 1 + +3.2.3 Coastal Eutrophication and “Dead Zones” +Excess phytoplankton NPP in coastal ecosystems can lead to accumulations o phytoplankton biomass and eutrophication. Anthropogenic N and P loading to estuarin and coastal marine ecosystems has more than doubled in the last 100 years (Seitzinge et al., 2010; Howarth et al., 2012),”” leading to a global spread of coastal eutrophicatio and associated increases in the number of oxygen-depleted “dead zones” (Duarte, 1995 Malone et al., 1999; Diaz and Rosenberg, 2008; Kemp et al., 2009), loss of sea grass bed (Dennison et al., 1993; Kemp et al., 2004; Schmidt et al., 2012), and increases in th occurrence of toxic phytoplankton blooms (see below). Current global trends in coasta eutrophication and the occurrence of “dead zones” and toxic algal events indicate tha phytoplankton NPP is increasing in many coastal ecosystems, a trend that is also likely t exacerbate future impacts of over-fishing, sea-level rise, and coastal development o ecosystem services (Dayton et al., 2005; Koch et al., 2009; Waycott et al., 2009). +3.2.4 Oxygen minimum zones (OMZs) +OMZs, which occur at midwater depths (200-1000 m) in association with easter boundary upwelling systems, are expanding globally as the solubility of dissolved O decreases and vertical stratification increases due to upper ocean warming (Chan et al. 2008; Capotondi et al., 2012; Bijma et al., 2013). Currently, the total surface area o ONZs is estimated to be ~30 x 10° km? (~8 per cent of the ocean’s surface area) with volume of ~10 x 10° km? (“0.1 per cent of the ocean’s volume). It is expected that th spatial extent of OMZs will continue to increase (Oschlies et al., 2008), a trend that i likely to affect nutrient cycles and fisheries — especially when combined with the sprea of coastal dead zones associated with coastal eutrophication. +3.2.5 Toxic Algal Blooms +Toxin-producing algae are a diverse group of phytoplankton species with only tw characteristics in common: (1) they harm people and ecosystems; and (2) thei initiation, development and dissipation are governed by species-specific populatio dynamics and oceanographic conditions (Cullen, 2008b). Negative impacts of algal toxin include illness and death in humans who consume contaminated fish and shellfish or ar exposed to toxins via direct contact (swimming, inhaling noxious aerosols); mas mortalities of wild and farmed fish, marine mammals and birds; and declines in th capacity of ecosystems to support goods and services (Cullen, 2008b; Walsh et al. 2008). Impacts associated with toxic algal blooms are global and appear to be increasin in severity and extent in coastal ecosystems as a consequence of anthropogeni nutrients, introductions of non-native toxic species with ballast water from ships, an climate-driven increases in water temperature and vertical stratification of the uppe ocean (Glibert et al., 2005; Glibert and Bouwman, 2012; Cullen, 2008b; Franks, 2008 Malone, 2008; Hallegraeff, 2010; Moore et al., 2008, Babin et al., 2008). +” Primarily due to the rapid rise in fertilizer use in agriculture, production of manure from farm animals domestic sewage, and atmospheric deposition associated with fossil-fuel combustion. +© 2016 United Nations 1 + +3.2.6 Nanoparticles +Nanoparticles have dimensions of 1-100 nm and are produced both naturally an anthropogenically. Of concern here are anthropogenic nanoparticles, such as titaniu dioxide (TiOz)*® and nanoplastics*®. Nanoparticulate TiO. is highly photoactive an generates reactive oxygen species (ROS) when exposed to ultraviolet radiation (UV) Consequently, TiO. has been used for antibacterial applications, such as wastewate treatment. It also has the potential to affect NPP. For example, it has been found tha ambient levels of UV from the sun can cause TiO2 nanoparticles suspended in seawate to kill phytoplankton, perhaps through the generation of ROS (Miller et al., 2012) Recent work has also highlighted the potential environmental impacts of microplastic (cf. Depledge et al., 2013; Wright et al., 2013). Experimental evidence suggests tha nanoplastics may reduce grazing pressure on phytoplankton and perturb nutrient cycles For example, Wegner et al., (2012) found that mussels (Mytilus edulis) exposed t nanoplastics consume less phytoplankton and grow slower than mussels that have no been exposed. In addition, microplastics contain persistent organic pollutants, and bot mathematical models and experimental data have demonstrated the transfer o pollutants from plastic to organisms (Teuten et al., 2009). +Understanding the ecotoxicology of anthropogenic nanoparticles in the marin environment is an important challenge, but as of this writing there is no clear consensu on environmental impacts in situ (cf. Handy et al., 2008). We know so little about th persistence and physical behaviour of anthropogenic nanoparticles in situ tha extrapolating experimental results, such as those given above, to the natural marin environment would be premature. We urgently need to develop the means to reliabl and routinely monitor nanoparticles of anthropogenic origin and their impacts o production and fate of phytoplankton biomass. A first step towards risk assessmen would be to establish and set limits based on their intrinsic toxicity to phytoplankto and the consumers of plankton biomass. The provision of such information is part of th mission of Working Group 40 of the Joint Group of Experts on the Scientific Aspects o Marine Environmental Protection (GESAMP). WG 40 was established to assess th sources, fate and effects of micro-plastics in the marine environment globally.”” +3.2.7 Ultraviolet Radiation and the Ozone Layer +The Sun emits ultraviolet radiation (UV, 400-700 nanometers), with UV-B (280-315 nm having a wide range of potentially harmful effects, including inhibition of primary +*® The world production of nanoparticulate TiO, is an order of magnitude greater than the next mos widely produced nanomaterial, ZnO. About 70 per cent of all pigments use TiO, and it is a commo ingredient in products such as sunscreen and food colouring. Titanium dioxide is therefore likely to ente estuaries and oceans, for example, from industrial discharge. +*® Plastic nanoparticles are released when plastic debris decomposes in seawater. Nanoparticles are als released from cosmetics and from clothes in the wash, and enter sewage systems where they ar discharged into the sea. +?° http://www.gesamp.org/work-programme/workgroups/working-group-40. +© 2016 United Nations 1 + +production by phytoplankton and cyanobacteria (Hader et al., 2007; Villar-Argaiz et al. 2009; Ha et al., 2012), changes in the structure and function of plankton communitie (Ferreyra et al., 2006; Hader et al., 2007; Fricke et al., 2011; Guidi et al., 2011; Santos e al., 2012a; Ha et al., 2014), and alterations of the N cycle (Goes et al., 1995; Jiang an Qiu, 2011). The ozone layer in the Earth’s stratosphere blocks most UV-B from reachin the ocean’s surface. Consequently, stratospheric ozone depletion since the 1970s ha been a concern, especially over the South Pole, where a so-called ozone hole ha developed.”* However, the average size of the ozone hole declined by ~2 per cen between 2006 and 2013 and appears to have stabilized, with variation from year to yea driven by changing meteorological conditions.”” It has even been predicted that ther will be a gradual recovery of ozone concentrations by ~2050 (Taalas et al., 2000). Give these observations and variations in the depths to which UV-B penetrates in the ocea (~1-10 m), a consensus on the magnitude of the ozone-depletion effect on NPP an nutrient cycling has yet to be reached. +4. Socioeconomic importance +Marine NPP supports a broad range of ecosystem services valued by society an required for sustainable development (Millennium Ecosystem Assessment, 2005; Wor et al., 2006; Conservation International, 2008; Perrings et al., 2010; Schlitzer et al., 2012 Malone et al., 2014b; Chapter 3 in this assessment). These include: +(1) food security through the production of harvestable fish, shellfish an macroalgae (Sherman and Hempel, 2009; Chassot et al., 2010; Barbier et al. 2011); +(2) climate regulation through carbon sequestration (Twilley et al., 1992; Cebrian 2002; Schlitzer et al., 2003; Duarte et al., 2005; Bouillon et al., 2008; Mitsch an Gosselink, 2008; Schneider et al., 2008; Subramaniam et al., 2008; Laffoley an Grimsditch, 2009; Nellemann et al., 2009; Chavez et al., 2011; Crooks et al. 2011; Henson et al., 2012); +(3) maintenance of water quality through nutrient recycling and water filtratio (Falkowski et al., 1998; Geider et al., 2001; Dayton et al., 2005; Howarth et al. 2011); +(4) protection from coastal erosion and flooding through the growth of macrophyt habitats (Danielsen et al., 2005; UNEP-WCMC, 2006; Davidson and Malone, +21 Ozone can be destroyed by reactions with by-products of man-made chemicals, such as chlorine fro chlorofluorocarbons (CFCs). Increases in the concentrations of these chemicals have led to ozon depletion. +2 http://www.nasa.gov/content/. +© 2016 United Nations 1 + +2006/2007; Braatz et al., 2007; Koch et al., 2009; Titus et al., 2009; Barbier et al. 2011), and +(5) development of biofuels and discovery of pharmaceuticals through th maintenance of biodiversity (Chynoweth et al., 2001; Orhan et al., 2006; Han e al., 2006; Yusuf, 2007; Negreanu-Pirjol et al., 2011; Vonthron-Sénécheau et al. 2011; Pereira et al., 2012; Sharma et al., 2012). +On a global scale, the value of these services in coastal marine and estuarine ecosystem has been estimated to be > 25 trillion United States dollars annually, making the coasta zone among the most economically valuable regions on Earth (Costanza et al., 1997 Martinez et al., 2007). The global loss of macrophyte ecosystems threatens the ocean’ capacity to sequester carbon from the atmosphere (climate control), suppor biodiversity (Part V of this Assessment) and living marine resources (Part IV of thi Assessment), maintain water quality, and protect against coastal erosion and floodin (Boesch and Turner, 1984; Dennison et al., 1993; Duarte, 1995; CENR, 2003; Scavia an Bricker, 2006; Davidson and Malone, 2006/07; Diaz and Rosenberg, 2008; MacKenzi and Dionne, 2008; Nellemann et al., 2009). Estimates of the value of these services b Koch et al., (2009) and Barbier et al., (2011) suggest that the socioeconomic impact o the degradation of marine macrophyte ecosystems is on the order of billions of U dollars per year. +5. Anthropogenic Impacts on Upper Ocean Plankton and Nutrient Cycles +5.1 Nitrogen loading +The rate of industrial Nitrogen gas (N2) fixation increased rapidly during the 20" centur and is now about equal to the rate of biological N2 fixation, resulting in a two- t threefold increase in the global inventory of Reactive nitrogen (N,) (Galloway et al. 2004; Howarth, 2008), a trend that has accelerated the global N cycle (Gruber an Galloway 2008). Today, anthropogenic N, inputs to surface waters via atmospheri deposition and river discharge are now roughly equivalent to marine N> fixation (Tabl 2) and are expected to exceed marine N> fixation in the near future as a result o increases in emissions from combustion of fossil fuels and use of synthetic fertilizers This trend is expected to continue (Duce et al., 2008; Seitzinger et al., 2010). +Atmospheric deposition of anthropogenic N, increased by an order of magnitude durin the 20" century to ~54 Tg N y1(80 per cent of total deposition), an amount that coul increase NPP by ~0.06 per cent. Estimates of anthropogenic emissions for 2030 indicat a 4-fold increase in total atmospheric N, deposition to the ocean and an 11-fold increas in AAN deposition (Duce et al., 2008). However, Lamarque et al., (2013) suggest tha oxidized Nr may decrease later this century because of increased control of the emissio of oxidized N compounds. At the same time, the geographic distribution of atmospheri deposition has also changed (Suntharalingam et al., 2012). In the late 1800s, +© 2016 United Nations 1 + +atmospheric deposition over most of the ocean is estimated to have been < 50 mg Nm y’. By 2000, deposition over large ocean areas exceeded 200 mg Nm~ y™ with intens deposition plumes (> 700 mg N m~ y”) extending downwind from Asia, India, North an South America, Europe and West Africa. Predictions for 2030 indicate similar patterns but with higher deposition rates extending farther offshore into the oligotrophic subtropical central gyres. Likewise, marine N2O production has increased compared t pre-industrial times downwind of continental population centres (in coastal and inlan seas by 15-30 per cent, in oligotrophic regions of the North Atlantic and Pacific by 5-2 per cent, and in the northern Indian Ocean by up to 50 per cent). These regiona patterns reflect a combination of high N, deposition and enhanced N2O production i suboxic zones. +The major pathway of anthropogenic N, loading to the oceans is river runoff Anthropogenic N, input to the coastal ocean via river discharge more than double during the 20" century due to increases in fossil-fuel combustion, discharges of huma and animal wastes, and the use of industrial fertilizers in coastal watersheds (Peierls e al., 1991; Galloway et al., 2004; Seitzinger et al., 2010). Riverine input of N, to th coastal ocean is correlated with human population density in and net anthropogeni inputs (NANI)?* to coastal watersheds (Howarth et al., 2012). NPP in coastal marine an estuarine ecosystems increases with increasing riverine inputs of N, (Nixon, 1992). Give predicted increases in population density in coastal watersheds and climate-drive changes in the hydrological cycle, global nutrient-export models predict that riverin inputs of N, to coastal waters will double again by 2050 (Seitzinger et al., 2010). In thi context, it is noteworthy that anthropogenic perturbations of the N-cycle caused b NANI already exceed the estimated “planetary boundary” (35 x 10° kg yr“) within whic sustainable development is possible (Rockstram et al., 2009). +Ocean warming and associated increases in vertical stratification are likely to exacerbat the effects of increases in NANI on phytoplankton NPP in coastal waters (Rabalais et al. 2009). As a consequence, excess NPP and the global extent of coastal eutrophication ar likely to continue increasing, especially in coastal waters near large watersheds population centres and areas of industrial agriculture (Kroeze and Seitzinger, 1998 Dayton et al., 2005; Seitzinger et al., 2005; UNESCO, 2008; Kemp et al., 2009; Rabalais e al., 2009; Sherman and Hempel, 2009). +5.2 Ocean warmin 5.2.1 Global impacts on NP Henson et al., (2013) used the results of six global biogeochemical models to project the +effects of upper ocean warming on the amplitude and timing of seasonal peaks in +3 Net anthropogenic nitrogen input (NANI) is the sum of synthetic N fertilizer used, N fixation associate with agricultural crops, atmospheric deposition of oxidized N, and the net movement of N into or out o the region in human food and animal feed. +© 2016 United Nations 2 + +phytoplankton NPP. Amplitude decreased by 1-2 per cent over most of the ocean except in the Arctic, where an increase of 1 per cent by 2100 is projected. These result are supported by the response of phytoplankton and zooplankton to global climate change projections carried out with the IPSL Earth System Model (Chust et al., 2014) Projected upper ocean warming by the turn of the century led to reductions i phytoplankton and zooplankton biomass of 6 per cent and 11 per cent, respectively Simulations suggest such declines are the predominant response over nearly 50 per cen of the ocean and prevail in the tropical and subtropical oceans while increasing trend prevail in the Arctic and Antarctic oceans. These results suggest that the capacity of th oceans to regulate climate through the biological carbon pump may decrease over th course of this century. The model runs also indicate that, on average, a 30-40 year tim series of observations will be needed to validate model results. +Regardless of the direction of global trends in NPP, climate change may be causing shift in phytoplankton community size spectra toward smaller cells which, if confirmed, wil have profound effects on the fate of NPP and nutrient cycling during this centur (Polovina and Woodworth, 2012). The size spectrum of phytoplankton communities i the upper ocean’s euphotic zone largely determines the trophic organization of pelagi ecosystems and, therefore, the efficiency with which NPP is channelled to higher trophi levels, is exported to the deep ocean, or is metabolized in the upper ocean (Malone 1980; Azam et al., 1983; Cushing, 1990; Kigrboe, 1993; Legendre and Rassoulzadegan 1996; Shurin et al., 2006; Pomeroy et al., 2007; Marafion, 2009; Barnes et al., 2010 Finkel et al., 2010; Suikkanen et al., 2013; and section 6.3.2). +In today’s ocean, the proportion of NPP accounted for by small phytoplankton (cell with an equivalent spherical diameter < 10 um) generally increases with increasin water temperature in the ocean (Atkinson et al., 2003; Daufresne et al., 2009; Marafion 2009; Huete-Ortega et al., 2010; Moran et al., 2010; Hilligsge et al., 2011) and wit increasing vertical stratification of the euphotic zone (Margalef, 1978; Malone, 1980 Kigrboe, 1993). Small cells also have a competitive advantage over large cells i nutrient-poor environments (Malone, 1980a; Chisholm, 1992; Kigrboe, 1993; Raven 1998; Marafion, 2009). Thus, as the upper ocean warms and becomes more stratified, i is likely that the small phytoplankton species will account for an increasingly larg fraction of NPP (Moran et al., 2010) resulting in increases in energy flow throug microbial food webs and decreases in fish stocks and organic carbon export to the dee sea (see section 6.1.1 and references therein). +This trend may be exacerbated by increases in temperature that are likely to stimulat plankton metabolism, enhancing both NPP and microbial respiration. Recent studie (Montoya and Raffaelli, 2010; Sarmento et al., 2010; Behrenfeld, 2011; Taucher an Oschlies, 2011; Taucher et al., 2012) suggest that predicted climate-driven increases i the temperature of the upper ocean will stimulate the NPP of smalle picophytoplankton cells (equivalent spherical diameter < 2um), despite predicte decreases in nutrient inputs to the euphotic zone from the deep sea in permanentl stratified regions of the ocean (e.g., the oligotrophic, subtropical central gyres). +© 2016 United Nations 2 + +However, if this does occur, it will not result in an increase in fishery production or i the ocean’s uptake of atmospheric CO2, because increases in picophytoplankton NP will be accompanied by equivalent increases in the respiratory release of CO. b bacterioplankton and other heterotrophic microbial consumers in the upper ocea (Sarmento et al., 2010; Behrenfeld, 2011). +5.2.2 Regional impacts on NPP +Regional trends in phytoplankton NPP are less controversial. The area of low NPP in th subtropical central gyres increased by 1-4 per cent yr’ from 1998 through 200 (Polovina et al., 2008; Vantrepotte and Mélin, 2009), a trend that is likely to continu through this century (Polovina et al., 2011). Decreasing NPP has been attributed t climate-driven (ocean warming) increases in vertical stratification and associate decreases in nutrient fluxes from deep water to the euphotic zone in the permanentl stratified subtropical gyres (Rost et al., 2008; Jang et al., 2011; Polovina et al., 2011 Capotondi et al., 2012; Moore et al., 2013). In the North Atlantic, upper ocean warmin and increases in stratification have been accompanied by decreasing NPP in water south of ~50°N, whereas warming and increases in stratification to the north have bee accompanied by increasing NPP (Richardson and Shoeman, 2004; Bode et al., 2011) These divergent responses to stratification reflect increases in the availability of sunligh in nutrient-rich, well-mixed subpolar waters and increases in nutrient limitation i nutrient-poor, permanently stratified”* subtropical waters (Richardson and Shoeman 2004; Steinacher et al., 2010; Bode et al., 2011; Capotondi et al., 2012). +Polar ecosystems are particularly sensitive to climate change (Smith et al., 2001 Anisimov et al., 2007; Bode et al., 2011; Doney et al., 2012; Engel et al., 2013), and th impacts of shrinking ice cover on NPP are expected to be especially significant in th Arctic Ocean (Wang and Overland, 2009). Loss of Arctic sea ice has accelerated in recen years (with a record low in 2012),”° a trend that is correlated with an increase in annua NPP by an average of 27.5 Tg C yr’ since 2003, with an overall increase of 20 per cen from 1998 to 2010 (Arrigo et al., 2008; Arrigo and van Dijken, 2011; Brown and Arrigo 2012). Of this increase, 30 per cent has been attributed to a decrease in the spatia extent of summer ice and 70 per cent to a longer growing season (the spring bloom i occurring earlier). The change in NPP is not spatially homogeneous. Positive trends ar most pronounced in seasonally ice-free regions, including the eastern Barents shelf Siberian shelves (Kara and east Siberian seas), western Mackenzie shelf, and the Berin Strait. NPP is expected to continue increasing during this century due to continued sea ice retreat and the associated increase in available sunlight. However, this trend may b short-lived if nitrate supplies from deep water are insufficient (Tremblay and Gagnon 2009). Neglecting the latter, Arrigo and van Dijken (2011) project a > 60 per cen increase in NPP for a summer ice-free Arctic using a linear extrapolation of the historical +“The permanent or main thermocline extends from ~50° N to ~50° S. North Atlantic Deep Water an Antarctic Bottom Water formation take place at higher latitudes 2s http://nsidc.org/arcticseaicenews//. +© 2016 United Nations 2 + +trend. Should these trends continue, additional loss of ice during Arctic spring coul boost NPP more than three-fold above 1998-2002 levels and alter marine ecosyste structure and the degree of pelagic-benthic coupling. However, predictions of futur trends in Arctic NPP are uncertain, given the possibility of nitrate limitatio (Vancoppenolle et al., 2013). Reducing uncertainty for both nitrate fields and rates o biogeochemical processes in the sea-ice zone should improve the skill of projecte changes in NPP needed to anticipate the impact of climate change on Arctic food web and the carbon cycle. +The coastal marine ecosystem of the West Antarctic Peninsula supports massive spring summer phytoplankton blooms upon which the production of Antarctic krill depends NPP associated with these blooms is correlated with the spatial and temporal extent o ice cover during the previous winter. Air temperatures over the West Antarcti Peninsula have warmed by 7°C since the 1970s, resulting in a 40 per cent decline i winter sea-ice cover and a decrease in phytoplankton NPP (Flores et al., 2012; Ducklo et al., 2013; Henley, 2013). Continued declines in the extent of winter sea-ice cover i likely to drive decadal-scale reductions in NPP and the production of krill, reducing th food supply for their predators (marine mammals, seabirds and people). +5.2.3 Distribution and abundance of toxic phytoplankton species +The socioeconomic impacts of toxic dinoflagellate species are increasing globally (Va Dolah, 2000; Glibert et al., 2005; Hoagland and Scatasta, 2006; Babin et al., 2008 UNESCO, 2012), and their distribution and abundance are sensitive indicators of th impacts of anthropogenic nutrient inputs and climate-driven increases in wate temperature and vertical stratification on ecosystem services (see section 6.3.2). +Alexandrium tamarense represents a group of species that cause paralytic shellfis poisoning (PSP) (Alexandrium catenella, A. fundyense, Pyrodinium bahamense an Gymnodinium catenatum) globally (Boesch et al., 1997). Since the 1970s, PSP episode have spread from coastal waters of Europe, North America and Japan to coastal water of South America, South Africa, Australia, the Pacific Islands, India, all of Asia and th Mediterranean (Lilly et al., 2007). Climate-driven shifts in the geographic ranges o Ceratium furca and Dinophysis spp. in the NE Atlantic have also occurred (Edwards et al. 2006), and the abundance of dinoflagellates in the North Sea have been positivel correlated with the North Atlantic Oscillation (NAO) and sea surface temperatur (Edwards et al., 2001). +5.2.4 Distribution and abundance of indicator zooplankton species +The distribution and abundance of calanoid copepods are also sensitive indicators o climate-driven increases in upper ocean temperature and basin-scale oscillations (Hay et al., 2005; Burkill and Reid, 2010; Edwards et al., 2010) including poleward shifts i species distributions (Beaugrand et al., 2002; Beaugrand et al., 2003; Cheung et al. 2010; Chust et al., 2014), decreases in size, and higher growth rates (e.g., Beaugrand e al., 2002; Richardson, 2008; Mackas and Beaugrand, 2010). There have also bee phenological changes, with the seasonal peak in abundance advancing to earlier in the +© 2016 United Nations 2 + +year for some species and being delayed for others (Edwards and Richardson, 2004 section 6.3.2). In the North Pacific, there is a strong correlation between sea-surfac temperature in the spring and the latitude at which subtropical species reach thei seasonal peak in abundance.”° Water temperature also influences the annual cycle o Neocalanus plumchrus biomass in the Northeast Pacific, where decadal-scale variation include a shift to an earlier occurrence of the seasonal biomass peak, as well as decrease in the duration of the bloom under warm ocean conditions (Mackas et al. 2007; Batten and Mackas, 2009). +The frequency and magnitude of gelatinous zooplankton blooms may be importan indicators of the status and performance of marine ecosystems (Hay, 2006; Graham e al., 2014). Both predators (medusa and ctenophores) and herbivores (tunicates) ca affect the fate of NPP (Pitt et al., 2009; Lebrato and Jones, 2011). Predators disrup metazoan food webs by consuming copepods and small fish (Richardson et al., 2009) Tunicates reduce the transfer of NPP to upper trophic levels and to the deep sea as thei gelatinous remains are degraded via microbial food webs (Lebrato and Jones, 2011). +Although, there is no evidence for an increase in the frequency and magnitude o gelatinous zooplankton on a global scale (Condon et al., 2012), decadal scale increase have been reported in several coastal marine ecosystems (Brodeur et al., 2002; Kideys 2002; Lynam et al., 2006; Uye, 2008; Licandro et al., 2010). A rigorous analysis of multi decadal (using a 1950 baseline) abundance data for 45 Large Marine Ecosystems, Brot et al., 2012 found that 28 (62 per cent) exhibited increasing trends while 3 (7 per cent exhibited decreasing trends. Thus, while increases of jellyfish populations may not b globally universal, they are both numerous and widespread. The most likely causes o these trends include ocean warming, overfishing, coastal eutrophication, habita modification, aquaculture, and introductions of non-indigenous gelatinous specie (Brotz et al., 2012; Purcell, 2012). While direct evidence is lacking for most of thes pressures, jellyfish tend to be most abundant in warm waters with low forage fis populations, and it is likely that ocean warming will provide a rising baseline o abundance leading to increases in the magnitude of jellyfish blooms and associate impacts on ecosystem services (Graham et al., 2014). +5.3 Ocean acidification +The oceans are becoming more acidic due to increases in uptake of atmospheric C (Calderia and Wickett, 2003; Calderia and Wickett, 2005; Doney et al., 2009; Beardall e al., 2009), and most of the upper ocean is projected to be undersaturated with respec to aragonite within 4-7 decades (Orr et al., 2005) with undersaturation expected t occur earliest at high latitudes (> 40°) and in upwelling systems where the aragonit saturation horizon is expected to shoal most rapidly (Feely et al., 2009, Gruber et al. 2009). These chemical changes in turn affect marine plankton via several mechanisms +6 http://www.pices.int/publications/pices_press/volume16/v16_n2/pp_19-21_CPR_f.pdf. +© 2016 United Nations 2 + +including the following: (1) decreases in the degree of aragonite saturation makes i harder for calcifying organisms (e.g., coccolithophores, foraminifera, and pteropods) t precipitate their mineral structures; (2) decreases in pH alters the bioavailability o essential algal nutrients such as iron and zinc; and (3) increases in CO, decrease th energy requirements for photosynthetic organisms to synthesize biomass. Suc biological effects are likely to perturb marine biogeochemical cycles including carbo export to the deep sea via the biological pump which may have a positive feedback o the buildup of CO in the upper ocean and atmosphere. However, assessments of th impacts of ocean acidification on NPP and nutrient cycling remain controversial and ar a subject of much research (cf., Delille et al., 2005; Doney et al., 2009; Shi et al., 2009 Shi et al.,2010; Shi et al., 2012; Moy et al., 2009; Kristy et al., 2010). For example increases in CO2 may stimulate Nz and carbon fixation by colonial cyanobacteria diazotrophs (Barcelos e Ramos et al., 2007). In addition, as the upper ocean warms, th geographic range of diazotrophs will expand. These effects may combine to enhance N fixation by as much as 35-65 per cent by the end of this century (Hutchins et al., 2009) It is noteworthy interesting that projected increases in N2 fixation are about the sam magnitude as increases in denitrification projected by Oschlies et al., (2008). Althoug both of these estimates have large uncertainties, if input and output fluxes accelerate a about the same rate, the ocean’s global inventory of N, would not change, whereas NP could increase (Sarmento et al., 2010; Behrenfeld, 2011). +In regard to macrophytes, photosynthetic rates of calcifying macroalgae do not appea to be stimulated by elevated CO, conditions, i.e., the majority of studies to date hav shown a decrease or no change in photosynthetic rates under elevated CO, condition (Hofmann and Bischof, 2014). On the other hand, there is clear evidence that ocea acidification (higher pCO2) stimulates seagrass NPP resulting in increases in above- an below-ground biomass suggesting that the capacity of seagrasses to sequester carbo may be significantly increased (Garrard and Beaumont, 2014). +5.4 Sea-level rise, coastal development and macrophyte NPP +Sea levels have increased globally since the 1970s, mainly as a result of global mean sea level rise due in part to anthropogenic warming causing ocean thermal expansion an glacier melting (Chapter 4 of this Assessment). Sea-level rise will not be unifor globally. Regional differences in sea-level trends will be related to changes in prevailin winds, ocean circulation, gravitational pull of polar ice sheets, and subsidence, so tha sea-level rise will exceed the global mean in some regions and will actually fall i others.” +To date, the global decline in macrophyte habitats has been primarily due to coasta development, artificially hardened shorelines, aquaculture operations, dredging an eutrophication. This will change with sea-level rise (Short and Neckles, 1999; Nicholls +27 http://tidesandcurrents.noaa.gov/sltrends//. +© 2016 United Nations 2 + +and Cazenave, 2010). Macrophyte habitats are projected to be negatively affected b sea-level rise and subsidence, especially where distributions are constrained on thei landward side by geomorphology and human activities along the shoreline (Pernetta 1993; Short and Neckles, 1999; Orth et al., 2006; Alongi, 2008; Gilman et al., 2008 Silliman et al., 2009; Waycott et al., 2009; Donato et al., 2011). Together, sea-level rise subsidence, coastal development and aquaculture operations are destroying mangrov forests, tidal marshes and seagrass beds at an alarming rate. The combination of sea level rise and the loss of these coastal habitats will decrease the capacity of coasta ecosystems to provide services, including climate regulation (carbon sequestration) protection against coastal flooding and erosion, and the capacity to support biodiversit and living marine resources. +5.5 Regions of special interes 5.5.1. Coastal river plumes +Increases in land-based anthropogenic inputs of N and P to coastal waters is drivin increases in annual phytoplankton NPP in estuaries and coastal marine ecosystems nea population centres and areas of industrial agriculture in large river basins (sections 6.2. and 6.2.2). This may lead to further increases in the spatial extent and/or number o coastal ecosystems experiencing eutrophication and oxygen-depleted dead zone associated with the coastal plumes of major river-coastal systems, including the Yangtz (E. China Sea), Mekong (S. China Sea), Niger (Gulf of Guinea), Nile (Mediterranean Sea) Parana (Atlantic Ocean), Mississippi (Gulf of Mexico), and Rhine (North Sea) (UNESCO 2012). +5.5.2 Polar waters and subtropical gyres +Ocean warming appears to be driving opposing trends in phytoplankton NPP in pola waters (interannual increases in NPP) and subtropical gyres (interannual decreases i NPP) and a global expansion of oxygen minimum zones associated with upwellin systems. Regions of special interest include the Arctic Ocean, coastal waters of th western Antarctic Peninsula, permanently stratified subtropical gyres of the Nort Pacific and North Atlantic, and major coastal upwelling centers (Cariaco Basin an California, Humboldt, Canary, Benguela and Somali Currents). +5.5.3 Subpolar waters +Early expressions of the impacts of ocean acidification on upper ocean plankton ar most likely to occur at high latitudes. Pteropods and foraminifera (dominated b Globigerina bulloides) are most abundant at high latitude (> 40°N) in surface waters of the North Atlantic (Barnard et al., 2004; Fraile et al., 2008 BednarSek et al., 2012), whereas the coccolithophore EF. huxleyi is most abundant in the +© 2016 United Nations 2 + +“Great Southern Coccolithophore Belt” of the Southern Ocean” and at high latitudes i the NE Atlantic (Barnard et al., 2004; Balch et al., 2011; Sadeghi et al., 2012). If th abundance of these functional groups declines in these regions, likely impacts will be t reduce the capacity of the oceans to take up CO2, export carbon to the deep sea, an support fisheries (Cooley et al., 2009). +6. Information needs +As shown above, anthropogenic nutrient-loading of coastal waters and climate-chang pressures on marine ecosystems (ocean warming and acidification, sea-level rise) ar driving changes in NPP and nutrient cycles that are affecting the provision of ecosyste services and, therefore, sustainable development. However, although changes i macrophyte NPP and their impacts are relatively well documented (and must continu to be), a consensus on the magnitude of changes and even the direction of change i phytoplankton NPP and upper ocean nutrient cycles has yet to be reached. +Documenting spatial patterns and temporal trends in NPP and nutrient cycles (and thei causes and socioeconomic consequences) will rely heavily on the accuracy an frequency with which changes in NPP and nutrient cycling can be detected over a broa range of scales (cf. deYoung et al., 2004; UNESCO, 2012; Mathis and Feeley, 2013) Given the importance of marine NPP and the species diversity of primary producers t sustaining ecosystem services, rapid detection of changes in time-space patterns o marine NPP and in the diversity of primary producers that contribute to NPP is a important dimension of the Regular Process” for global reporting and assessment o the state of the marine environment, including socioeconomic aspects. +Data requirements for the Regular Process have been used to help guide th development of the Global Ocean Observing System and an implementation strategy fo its coastal module (UNESCO, 2012; Malone et al., 2014a; Malone et al., 2014b). Th essential variables required to compute key indicators of ecosystem health includ species richness, chlorophyll-a, dissolved N,, and dissolved BAP (UNESCO, 2012). Routin and sustained measurements of these variables over a range of temporal and spatia scales are required for rapid and timely detection of changes in NPP and nutrient cycle and the impacts of these changes on ecosystem services on regional (e.g., Large Marin Ecosystems) to global scales. Although satellite imagery, limited in situ measurement and numerical models are making it possible to detect interannual and decadal change in NPP on these scales, the same cannot be said for observations of species richness an nutrient distributions (UNESCO, 2012). +*8 The belt is centered around the sub-Antarctic front and has a spatial extent of 88 x 106 km? (~25 pe cent of the global ocean) *° http://www.un.org/Depts/los/global_reporting/global_reporting.htm. +© 2016 United Nations 2 + +6.1 Net primary production +Sustained observations of chlorophyll, irradiance and temperature fields are require for model-based estimates of phytoplankton NPP (see section 6.1.2). An integrate approach using long term data streams from both remote sensing and frequent in sit observations is needed to capture the dynamics of marine phytoplankton NPP and t detect decadal trends. Remote sensing provides a cost-effective means to observ physical and biological variables synoptically in time and space with sufficient resolutio to elucidate linkages between climate-driven changes in the NPP of ecosystems and th dynamic relationship between phytoplankton NPP and the provision of ecosyste services (Platt et al., 2008; Forget et al., 2009). For details on requirements, advantage and limitations of satellite-based remote sensing of ocean colour, see IOCCG (1998) Sathyendranath (2000), and UNESCO (2006, 2012). +Two related activities, both contributions to the Global Ocean Observing System provide the core of an integrated observing system needed to provide data required t assess the state of the marine environment in terms of both time-space variations i phytoplankton NPP and the impacts of these variations on ecosystem services: th Chlorophyll Global Integrated Network (ChloroGIN)*° (Sathyendranath et al., 2010) an Societal Applications in Fisheries and Aquaculture using Remotely-Sensed Imager (SAFARI) (Forget et al., 2010). FARO (Fisheries Applications of Remotely Sensed Ocea Colour) has recently been initiated to coordinate the development of ChloroGIN an SAFARI for the provision of ocean-colour data and data products for use in fisherie research and ecosystem-based management of living marine resources.** Likewise, th GEO Biodiversity Observation Network, the Global Biodiversity Information Facilit (GBIF), and the Ocean Biogeographical Information System (UNESCO, 2012) provid data and information on the species richness of marine primary producers. +6.2 Nitrogen and phosphorus cycles +The N cycle is more dynamic* and less well understood than previously though (Codispoti et al., 2001; Capone and Knapp, 2007; Zehr and Kudela, 2011; Landolfi et al. 2013; Voss et al., 2013). Major impediments to detecting and understanding decada changes in the marine N cycle are: current uncertainties about the rate (undersampling); distribution and coupling of sources and sinks; sensitivity of N2 fixation denitrification, and anammox to anthropogenic inputs of N,; and changes in the marin environment associated with climate change (warming and increases in stratification o the upper ocean, ocean acidification, oxygen depletion, and sea-level rise). +°° http://www.chlorogin.org/. +3 http://www.faro-project.org/index.html. +2 Estimates of turnover times of N, have decreased from 10,000 years to 1,500 years (Codispoti et al. 2001). +© 2016 United Nations 2 + +Quantifying inputs of N and P to coastal ecosystems and the open ocean requires network of coordinated and sustained observations on local to global scales. Fo atmospheric deposition, monitoring should focus on regions that have intens deposition plumes downwind of major population centres in West Africa, East Asia Europe, India, North and South America (section 6.2.1 and Schulz et al., 2012). This is major goal of the SOLAS programme **. Shipboard time-series observations o biogeochemical variables that are being established globally** should provide depositio data for these plumes. For riverine inputs, rivers that are part of the Global Terrestria Network for River Discharge (GTN-R)*° and that represent a broad range of volum discharges and catchment-basin population densities are high priorities for monitorin land-based inputs and associated land-cover/land-use practices in their watershed (UNESCO, 2012). +All global ocean biogeochemistry models require oceanographic data on physical an chemical variables, including temperature, salinity, mixed-layer depth, and th concentration of macro-nutrients (N, P, Si) (Le Quéré et al., 2010). Over the last decade autonomous technologies for measuring essential physical variables (includin temperature, salinity and mixed-layer depth) have revolutionized our ability to observ the sea surface and the ocean’s interior. By integrating data from both remote sensin (satellite-based sensors and land-based HF radar) and in situ measurements (from ship of opportunity, research vessels and automated moorings, profiling floats, gliders surface drifters and large pelagic predators), observations of atmospheric and uppe ocean geophysics are now made continuously in four dimensions; data are transmitte to data assembly centers in near-real time via satellites, fiber-optic cables, and th internet; and predictions (nowcasts and forecasts) of atmospheric and upper ocea “weather” are made routinely using data assimilation techniques and couple atmospheric-hydrodynamic models (Hall et al., 2010). +Over the last decade, autonomous technologies have revolutionized our ability t measure nitrate, nitrite, ammonium and reactive phosphate in situ (Johnson and Coletti 2002; ACT, 2003; Sakamoto et al., 2004; Adornato et al., 2010). Efforts are als underway to expand sampling programmes such Repeat Hyrdrography (Hood 2009) Argo (Rudnick et al., 2004; Testor et al., 2010), and OceanSites** to incorporate in sit nutrient sensors. +33 http://www.solas-int.org/. +34 e.g., For example, http://www.unesco.org/new/en/natural-sciences/ioc-oceans/sections-and programmes/ocean-sciences/biogeochemical-time-series/. +35 http://www.fao.org/gtos/gt-netRIV.html; http://gtn-r.bafg.de http://www.bafg.de/GRDC/EN/Home/homepage_node.html. +36 http://www.whoi.edu/virtual/oceansites/ +© 2016 United Nations 2 + +6.3 Plankton species diversity +Sustaining marine species richness” is the single most important indicator of th capacity of ecosystems to support services valued by society (Worm et al., 2006). biodiversity observation network (GEO BON)*® has been established to documen changes in species biodiversity, and the Ocean Biogeographic Information Syste (OBIS)*? documents the species diversity, distribution and abundance of life in th oceans. Both are contributions to GEOSS.”” A set of sentinel sites should be targeted fo sustained observations of species richness including Large Marine Ecosystems and th emerging network of marine protected areas that is nested within them (Malone et al. 2014a). As a group, these sites represent a broad range of species diversity, climate related changes in the marine environment, and anthropogenic nutrient inputs. Here w underscore the importance of rapid detection of changes in plankton diversity and earl warnings of impacts on marine ecosystem services. +References +Adornato, L., Cardenas-Valencia, A., Kaltenbacher, E., Byrne, R.H., Daly, K., and other (2010). In situ nutrient sensors for ocean observing systems. In Proceedings o OceanObs’09: Sustained Ocean Observations and Information for Society (Vol. 2) Venice, Italy, 21-25 September 2009. Hall, J., Harrison, D.E., & Stammer, D., eds ESA Publication WPP-306, doi:10.5270/OceanObs09.cwp.01. +ACT (Alliance for Coastal Technologies) (2003). State of Technology in the Developmen and Application of Nutrient Sensors; ACT (2007). Recent Developments in /n Sit Nutrient Sensors: Applications and Future Directions, ACT 06-08 (UMCES CBL 07 048). +Agawin, N.S.R., Duarte, C.M., and Agusti, S. (2000). Nutrient and temperature control o the contribution of picoplankton to phytoplankton biomass and production Limnology and Oceanography, 45(3): 591-600. +Alongi, D. (2008). Mangrove forests: Resilience, protection from tsunamis, an responses to global climate change. Estuarine Coastal and Shelf Science, 76, 1-13. +37 Species richness is an unweighted list of species present in an ecosystem that is especially important t monitor because it is the simplest indicator of species diversity and it does not discount rare species tha are often the primary concern. +38 https://www.earthobservations.org/geobon.shtml +39 http://www. iobis.org/ +“0 www.earthobservations.org/geobon.shtml +© 2016 United Nations 3 + +Altabet, M.A. (2007). Constraints on oceanic N balance/imbalance from sedimentary *° records, Biogeosciences, 4: 75-86. +Amarasinghe, M.D., and Balasubramaniam, S. (1992). Net primary productivity of tw mangrove forests stands on the northwestern coast of Sri Lanka. Hydrobiologia 247: 37-47. +Anisimov, O.A., Vaughan, D.G., Callaghan, T.V., Furgal, C., Marchant, H., Prowse, T.D. Vilhjalmsson, H., and Walsh, J.E. (2007). Polar regions (Arctic and Antarctic) Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution o Working Group II to the Fourth Assessment Report of the Intergovernmenta Panel on Climate Change. Parry, M.L., Canziani, O.F., Palutikof, J.P., van de Linden, P.J., and Hanson, C.E., eds. Cambridge University Press, Cambridge, 653 685. +Antoine, D., and Morel, A. (1996). Oceanic primary production. 1. Adaptation of spectral light-photosynthesis model in view of application to satellite chlorophyl observations. Global Biogeochemical Cycles, 10 (1): 43-55. +Appeltans, W., Ahyong, S.T., Anderson, G., Angel, M.V., Artois, T., et al. (2012). Th magnitude of global marine species diversity. Current Biology, 22: 2189 — 2202. +Arrigo, K.R. (2005). Marine microorganisms and global nutrient cycles. Nature, 437, 349 355. +Arrigo, K.R. and van Dijken, G.L. (2011). Secular trends in Arctic Ocean net primar production, Journal of Geophysical Research, 116, CO9011 doi:10.1029/2011JC007151. +Arrigo, K.R., van Dijken, G.L., and Pabi, S. (2008). Impact of a shrinking Arctic ice cove on marine primary production, Geophysical Research Letters, 35(19), L19603. +Atkinson, D., Ciotti, B.J., and Montagnes, D.J.S. (2003). Protists decrease in size linearl with temperature. Proceedings of the Royal Society of London B, 270: 2605-2611. +Azam, F., Fenchel, T., Field, J. G., Gray, J. S., Meyer-Reil, L. A., Thingstad, F. (1983). Th Ecological Role of Water-Column Microbes in the Sea. Marine Ecology Progres Series, 10: 257-263. +Babin, M., Roesler, C.S., and Cullen, J.J., eds. (2008). Real-Time Coastal Observin Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory Instrumentation and Modelling. Monographs on Oceanographic Methodolog Series, UNESCO. +Balch, W.M., Drapeau, D.T., Bowler, B.C., Lyczskowski, E., Booth, S., and Alley, D. (2011) The contribution of coccolithophores to the optical and inorganic carbo budgets during the Southern Ocean Gas Exchange Experiment: New evidence i support of the “Great Calcite Belt” hypothesis, Journal of Geophysical Research 116, COOFO6, doi:10.1029/2011JC006941. +Balch, W.M., and Utgoff, P.E. (2009). Potential interactions among ocean acidification, +© 2016 United Nations 3 + +coccolithophores, and the optical properties of seawater. Oceanography, 22 (4) 146-159. +Barber, R.T., Sanderson, M.P., Lindley, S.T., Chai, F., Newton, J., Trees, C.C., Foley, D.G. and Chavez, F.P. (1996). Primary productivity and its regulation in the equatoria Pacific during and following the 1991-92 El Nifio. Deep-Sea Research II, 43: 933 969. +Barbier, E.B., Hacker, S.D., Kennedy, C., Koch, E.W., Stier, A.C., and Silliman, B.R. (2011) The value of estuarine and coastal ecosystem services. Ecological Monographs 81: 169-193 +Barcelos Ramos, J., Biswas, H., Schulz, K.G., LaRoche, J., and Riebesel, U. (2007). Effect o rising atmospheric carbon dioxide on the marine nitrogen fixer Trichodesmium Global Biogeochemical Cycles, 21: GB2028, doi:10.1029/2006GB002898. +Barnard, R., Batten, S.D., Beaugrand, G. et al. (2004). Continuous Plankton Records Plankton atlas of the North Atlantic Ocean (1958 — 1999). II. Biogeograhica charts. Marine Ecology Progress Series, Supplement: 11-75. +Barnes, C., Maxwell, D., Reuman, D.C., and Jennings, S. (2010). Global patterns i predator-prey size relationships reveal size dependency of trophic transfe efficiency. Ecology, 91:222-232. +Barnes, C., lrigoien, X., De Oliveira, J.A.A., Maxwell, D., and Jennings, S. (2011) Predicting marine phytoplankton community size structure from empirica relationships with remotely sensed variables. Journal of Plankton Research, 33 13-24. +Barton, A.D., Dutkiewicz, S., Flierl, G., Bragg, J., Follows, M.J. (2010). Patterns of diversit in marine phytoplankton. Science, 327, 1509-1511. +Batten, S.D., and Mackas, D.L. (2009). Shortened duration of the annual Neocalanu plumchrus biomass peak in the Northeast Pacific. Marine Ecology Progress Series 393: 189-198. +Beardall, J., Stojkovic, S., and Larsen S. (2009). Living in a high CO world: impacts o global climate change on marine phytoplankton. Plant Ecology & Diversity, 2 191-205. +Beaugrand, G., K.M. Brander, J.A. Lindley, S. Souissi, and P.C. Reid. (2003). Plankto effect on cod recruitment in the North Sea. Nature, 426: 661-664. +Beaugrand, G., Reid, P.C., Ibafiez, F., Lindley, J.A., and Edwards, M. (2002) Reorganization of the North Atlantic Marine Copepod Biodiversity and Climate Science, 296: 1692-1694. +BednarSek, N., Tarling, G.A., Bakker, D.C.E., Fielding, S., Jones, E.M., Venables, H.J. Ward, P., Kuzirian, A., Lézé, B., Feely, R.A., Murphy, E.J. (2012). Extensiv dissolution of live pteropods in the Southern Ocean. Nature, 5: 881-885. +© 2016 United Nations 3 + +Behrenfeld, M.J. (2011). Biology: Uncertain future for marine algae. Nature Climat Change, 1: 33-34. +Behrenfeld, M.J., O’Malley, R.T., Siegel, D.A., McClain, C.R., et al. (2006). Climate-drive trends in contemporary ocean productivity. Nature 444: 752-755. +Behrenfeld, M.J., and Falkowski, P.G. (1997). Photosynthetic Rates Derived fro Satellite-Based Chlorophyll Concentration. Limnology and Oceanography, 42: 1 20. +Benitez-Nelson, C.R. (2000). The biogeochemical cycling of phosphorus in marin systems. Earth-Science Review, 51(1-4): 109-135. +Beusen, A.H.W., Slomp, C.P., and Bouwman, A.F. (2013). Global land—ocean linkage direct inputs of nitrogen to coastal waters via submarine groundwater discharge Environmental Research Letters, 8: 034035 (doi:10.1088/1748-9326/8/3/034035). +Bidigare, R.R., Chai, F., Landry, M.R., Lukas, R., Hannides, C.C.S., Christensen, S.J., Karl D.M., Shi, L., and Chao. Y. (2009). Subtropical ocean ecosystem structure change forced by North Pacific climate variations. Journal of Plankton Research, 31 (10) 1131-39. +Bijma, J., Portner, H-O. Yesson, C., and Rogers, A.D. (2013). Climate change and th oceans — What does the future hold? Marine Pollution Bulletin. 74 (2): 495-505. +Bissinger, J.E., Montagnes, D.J.S., Sharples, J., and Atkinson, D. (2008). Predicting marin phytoplankton maximum growth rates from temperature: Improving on th Eppley curve using quantile regression. Limnology and Oceanography, 53: 487 93. +Bittaker, H.F., and Iverson R.L. (1976). Thalassia testudinum productivity: A fiel comparison of measurement methods. Marine Biology, 37: 39-46. +Blanchard, J.L., Jennings, S., Holmes, R., Harle, J., Merino, G., Allen, J. |., Holt, J., Dulvy N K., and Barange, M. (2012). Potential consequences of climate change fo primary production and fish production in large marine ecosystems Philosophical transactions of the Royal Society of London, Series B, Biologica Sciences, 367: 2979-89. +Bode, A., Hare, J., Li, W.K.W., Moran, X.A.G., and Valdés L. (2011). Chlorophyll an primary production in the North Atlantic. In: Reid, P.C., and Valdés, L., eds. ICE Status Report on Climate Change in the North Atlantic. International Council fo the Exploration of the Sea, Copenhagen, pp. 77-102. +Boesch, D.F., and Turner, R.E. (1984). Dependence of Fishery Species on Salt Marshes The Role of Food and Refuge. Estuaries and Coasts, 7:460-468. +Boesch, D.F., Anderson, D.M., Horner, R.A., Shumway, S.E., Tester, P.A., an Whitledge, T.E. (1997). Harmful Algal Blooms in Coastal Waters: Options fo Prevention, Control and Mitigation. Science for Solutions, NOAA Coastal Ocean +© 2016 United Nations 3 + +Program Decision and Analysis Series, No. 10, NOAA Coastal Ocean Office, Silve Spring, MD, 46 pp. + appendix. +Bouillon, S., Borges, A.V., Castafieda-Moya, E., Diele, K., Dittmar, T., Duke, N.C. Kristensen, E., Lee, S.Y, Marchand, C., Middelburg, J.J., Rivera-Monroy, V.H. Smith III, T.J, and Twilley, R.R. (2008). Mangrove production and carbon sinks: revision of global budget estimates. Global Biogeochemical Cycles, 22(2) doi:10.1029/2007GB003052. +Boyce, D.G., Dowd, M., Lewis, M.R., Worm, B. (2014). Estimating global chlorophyl changes over the past century. Progress in Oceanography, 122: 163-177. +Boyce D.G., Lewis, M.R., and Worm, B. (2010). Global phytoplankton decline over th past century. Nature, 466: 591-596. +Boyer, E.W., Howarth, R.W., Galloway, J.N., et al. (2006). Riverine nitrogen export fro the continents to the coasts. Global Biogeochemical Cycle, 20: GB1S91 doi:10.1029/2005GB002537. +Braatz, S., Fortuna, S., Broadhead, J., and Leslie, R. (2007). Coastal protection in th aftermath of the Indian Ocean tsunami: What role for forests and trees? Ra publication, 2007/07, Proceedings of the Regional Technical Workshop, Kha Lak, Thailand, 28-31 August, 2006. +Brodeur, R.D., Sugisaki, H., Hunt, G.L. Jr., (2002). Increases in jellyfish biomass in th Bering Sea: implications for the ecosystem. Marine Ecology Progress Series, 233, 89 103. +Brotz, L., Cheung, W.W.L., Kleisner, K., Pakhomov, E., Pauly, D. (2012). Increasin jellyfish populations: trends in Large Marine Ecosystems. Hydrobiologia, 690: 3-20. +Brown, Z.W., and Arrigo, K.R. (2012). Contrasting trends in sea ice and primar production in the Bering Sea and Arctic Ocean /CES Journal of Marine Science, 69 1180-1193. +Buitenhuis, E.T., Li, W.K.W., Vaulot, D., Lomas, M.W., Landry, M.R., Partensky, F. Karl, D.M., Ulloa, O., Campbell, L., Jacquet, S., Lantoine, F., Chavez, F., Macias, D. Gosselin, M., and McManus, G.B., (2012). Picophytoplankton biomas distribution in the global ocean. Earth System Science Data, 4: 37-46. +Buesseler, K.O., Antia, A.N., Chen, M., Fowler, S.W., Gardner, W.D., Gustafsson, O. Harada, K., Michaels, A.F., van der Loeff, M.R., Sarin, M., Steinberg, D.K., an Trull, T. (2007). An assessment of the use of sediment traps for estimating uppe ocean particle fluxes. Journal of Marine Research, 65: 345-416. +Burkill, P.H., and Reid, P.C. (2010). Plankton biodiversity of the North Atlantic: changin patterns revealed by the Continuous Plankton Recorder survey. In: Proceeding of OceanObs’09: Sustained Ocean Observations and Information for Society, (Vol 1). Hall, J., Harrison, D.E., and Stammer, D., eds. ESA Publication WPP-306 doi:10.5270/OceanObs09.pp.09. +© 2016 United Nations 3 + +Calbet, A. (2008). The trophic roles of microzooplankton in marine systems. ICES Journa of Marine Science, 65: 325-331. +Caldeira, K., and Wickett, M.E. (2003). Oceanography: Anthropogenic carbon and ocea pH. Nature, 425: 365-365. +Caldeira K., and Wickett, M.E. (2005). Ocean model predictions of chemistry change from carbon dioxide emissions to the atmosphere and ocean. Journal o Geophysical Research, 110:C09S04, doi:10.1029/2004JC002671. +Capone, D.G. (2008). The Marine Nitrogen Cycle. Microbe, 3 (4): 186 — 192. +Capone, D.G., and Knapp, A.N. (2007). Oceanography: A marine nitrogen cycle fix Nature, 445: 159-160. +Capone, D.G., Bronk, D.A., Mulholland, M.R., and Carpenter, E.J. (2008). Nitrogen in th Marine Environment. Elsevier Inc., Amsterdam, 1729 pp. pp. +Capotondi, A., Alexander, M.A., Bond, N.A., Curchitser, E.N., and Scott, J.D. (2012) Enhanced upper ocean stratification with climate change in the CMIP3 models Journal of Geophysical Research, 117, C04031, doi:10.1029/2011JC007409. +Carr, M-E., Friedrichs, M.A.M., et al. (2006). A comparison of global estimates of marin primary production from ocean color. Deep-Sea Research Part II, 53: 741-770. +Cebrian, J. (2002). Variability and control of carbon consumption, export, an accumulation in marine communities. Limnology and Oceanography, Inc., 47(1) 11-22. +CENR. (2003). An Assessment of Coastal Hypoxia and Eutrophication in U.S. Waters Washington DC: National Science and Technology Council, Committee o Environment and Natural Resources. +Chan, F., Barth, J.A., Lubchenco, J., et al. (2008). Emergence of Anoxia in the Californi Current Large Marine Ecosystem. Science, 319: 920. +Charpy-Roubaud, C., and Sournia, A. (1990). The comparative estimation o phytoplanktonic, microphytobenthic and macrophytobenthic primary productio in the oceans. Marine Microbial Food Webs, 4 (1): 31-57. +Chassot, E., Melin, F., Le Pape, O., and Gascuel, D. (2007). Bottom-up control regulate fisheries production at the scale of eco-regions in European seas. Marine Ecolog Progress Series, 343: 45-55. +Chassot, E., Bonhommeau, S., Dulvy, N.K., Mélin, F., Watson, R., Gascuel, D., an Le Pape. O. (2010). Global marine primary production constrains fisherie catches. Ecology Letters, 13(4): 495 — 505. +Chavez, F.P., Messié, M., and Pennington, J.T. (2011). Marine Primary Production i Relation to Climate Variability and Change. Annual Reviews Marine Science, 3 227-260. +© 2016 United Nations 3 + +Chaalali A., Beaugrand, G., Raybaud, V., Goberville, E., David, V., et al. (2013). Climati Facilitation of the Colonization of an Estuary by Acartia tonsa. PLoS ONE 8&(9) e74531. doi:10.1371/journal.pone.0074531 +Cheung, W., Lam, V., Sarmieno, J.L., Kearney, K., Watson, R., Zeller, D., Pauly, D. (2010) Large scale redistribution of maximum fisheries catch potential in the globa ocean unser climate change. Global Change Biology., 16: 24-35. DOI 10.1111/j.1365-2486.2009.01995.x +Chisholm, S.W. (1992). Phytoplankton size, p. 213-237. In: Falkowski, P.G., an Woodhead, A.D., (eds.), Primary Productivity and Biogeochemical Cycles in th Sea. Plenum Press. +Church, J.A., and White, N.J. (2006). A 20th century acceleration in global sea-level rise Geophysical Research Letters, 33: L01602. +Chust, G., Allen, J. 1., Bopp, L., Schrum, C., Holt, J., Tsiaras, K., Zavatarelli, M., Chifflet, M. Cannaby, H., Dadou, |., Daewel, U., Wakelin, S. L., Machu, E., Pushpadas, D. Butenschon, M., Artioli, Y., Petihakis, G., Smith, C., Gargon, V., Goubanova, K., +Le Vu, B., Fach, B. A., Salihoglu, B., Clementi, E. and Irigoien, X. (2014). Biomas changes and trophic amplification of plankton in a warmer ocean. Global Chang Biology, 20: 2124-2139. doi: 10.1111/gcb.12562. +Chynoweth, D.P., Owens, J.M., and Legrand, R. (2001). Renewable methane fro anaerobic digestion of biomass. Renewable Energy, 22:1-8. +Cleland, E.E. (2011). Biodiversity and Ecosystem Stability. Nature Education Knowledge 3(10):14 +Cloern, J.E. (1987). Turbidity as a control on phytoplankton biomass and productivity i estuaries. Continental Shelf Research, 7: 1367-1381. +Cloern, J.E. (2001). Review. Our evolving conceptual model of the coastal eutrophicatio problem. Marine Ecology Progress Series, 210: 223-253. +Cloern, J.E., and Jassby, A.D. (2010). Patterns and Scales of Phytoplankton Variability i Estuarine—Coastal Ecosystems. Estuaries and Coasts, 33:230-241. +Cloern, J.E., Foster, S.Q., and Kleckner, A.E. (2013). Review: phytoplankton primar production in the world’s estuarine-coastal ecosystems. Biogeoscience Discussion, 10: 17725-17783. +Codispoti, L.A., Brandes, J.A., Christensen, J.P., Devol, A.H., Naqvi, S.W.A., Paerl, H.W. and Yoshinari, T. (2001). The oceanic fixed nitrogen and nitrous oxid budgets:Moving targets as we enter the anthropocene? Scientia Marina 65(Suppl. 2): 85-105. +Compton, J.S., Mallinson, D., Glenn, C.R., Filippelli, G.M., Follmi, K.B., Shields-Zhou, G.A. and Zanin, Y. (2000). Variations in the global phosphorus cycle, in Marin Authigenesis: From Global to Microbial, Glenn, C.R., Prévol-Lucas, L., and Lucas J., eds. SEPM Special Publication, 66, 21-33. +© 2016 United Nations 3 + +Condon, R. H., Graham, W.M., Duarte, C.M., Pitt, K.A., Lucas, C.H., Haddock, S.H.D. Sutherland, K.R., Robinson, K.L., Dawson, M.N., Decker, M.B., and others. (2012) Questioning the Rise of Gelatinous Zooplankton in the World’s Oceans. BioScience 62: 160-169. +Conkright, M.E., and Gregg, W.W. (2003). Comparison of global chlorophyl climatologies: In situ, CZCS, blended in situ-CZCS and SeaWiFS. Internationa Journal of Remote Sensing, 24 (5): 969-991. +Conservation International. (2008). Economic Values of Coral Reefs, Mangroves, an Seagrasses: A Global Compilation. Center for Applied Biodiversity Science Arlington, VA. +Cooley, S.R., Kite-Powell, H.L., and Doney, S.C. (2009). Ocean acidification’s potential t alter global marine ecosystem services. Oceanography, 22: 172-181. +Corten, A., and Lindley, J.A. (2003). The use of CPR data in fisheries research. Progress i Oceanography, 58: 285-300. +Costanza, R., d’Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., Limburg, K. Naeem, S., O’Neill, R.V., Paruelo, J., Raskin, R.G., Suttonk, P., an van den Belt, M. (1997). The value of the world’s ecosystem services and natura capital. Nature, 387: 253-260. +Cousens, R. (1984). Estimation of Annual Production by the Intertidal Brown Alg Ascophyllum nodosum (L.) Le Jolis. Botanica Marina, 27: 217-227. +Craig, H. (1957). The Natural Distribution of Radiocarbon and the Exchange Time o Carbon Dioxide between Atmosphere and Sea. Tellus, 9(1): 1-17. +Crooks, S., Herr, D., Tamelander, J., Laffoley, D., and Vandever, J. (2011). Mitigatin Climate Change through Restoration and Management of Coastal Wetlands an Near-shore Marine Ecosystems: Challenges and Opportunities. Environmen Department Paper 121, World Bank, Washington DC. +Cullen, J.J. (2008a). Primary production methods. pp. 578-584, In Steele, J.H., Turekian K.K., and Thorpe, S.A., eds. Encyclopedia of Ocean Science, Elsevier, ISBN: 978-0 12-374473-9. +Cullen, J.J. (2008b). Observation and prediction of harmful algal blooms, In Babin, M. Roesler, C.S., and Cullen, J.J., eds. Real-Time Coastal Observing Systems fo Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory, Instrumentatio and Modelling. Monographs on Oceanographic Methodology Series, UNESCO pp. 1-41. +Cullen, J.J., Davis, R.F., and Huot, Y. (2012). Spectral model of depth-integrated wate column photosynthesis and its inhibition by ultraviolet radiation. Globa Biolgeochemical Cycles, 26, GB1011, doi:10.1029/2010GB003914. +© 2016 United Nations 3 + +Curcé, A., Ibafiez, C., Day, J.W., and Prat, N. (2002). Net primary production an decomposition of salt marshes of the Ebre delta (Catalonia, Spain). Estuaries an Coasts, 25(3): 309-324. +Cushing, D.H. (1990). Plankton production and year-class strength in fish populations: a update of the match/mismatch hypothesis. Advances in Marine Biology, 26: 249 293. +Dame, R.F., and Kenny, D. (1986). Variability of Spartina alterniflora primary productio in the euhaline North Inlet estuary. Marine Ecology - Progress Series, 32: 71-80. +Danielsen, F., SArensen, M.K., Olwig, M.F., Selvam, V., Parish, F., Burgess, N.D. Hiraishi, T., Karunagaran, V.M., Rasmussen, M.S., Hansen, L.B., Quarto, A., an Suryadiputra, N. (2005). The Asian tsunami: A protective role for coasta vegetation. Science, 310: 643. +Daufresne, M., Lengfellner, K., and Sommer, U. (2009). Global warming benefits th small in aquatic ecosystems. Proceedings of the National Academy of Sciences USA 106: 12788-12793 +Dave, A.C., and Lozier, M.S. (2013). Examining the global record of interannual variabilit in stratification and marine productivity in the low-latitude and mid-latitud ocean. Journal of Geophysical Research. Oceans, 118, 3114-3127 doi:10.1002/jgrc.20224. +Davidson, M., and Malone, T.C., eds. (2006/2007). Stemming the tide of coasta disasters. Marine Technology Society Journal, 40: 1-125. +Day, J.W., Jr., Coronado-Molina, C., Vera-Herrera, F.R., Twilley, R., Rivera-Monroy, V.H. Alvarez-Guillen, H., Day, R., and Conner, W. (1996). A 7 year record of above ground net primary production in a southeastern Mexican mangrove forest Aquatic Botany, 55: 39-60. +Dayton, P., Curran, S., Kitchingman, A., Wilson, M., Catenazzi, A., Restrepo, J. Birkeland, C., Blaber, S., Saifullah, S., Branch, G., Boersma, D., Nixon, S. Dugan, P., Davidson, N., and V6rdésmarty, C. (2005). Coastal ecosystems In Ecosystems and Human Well-being: Current State and Trends: Findings of th Condition and Trends Working Group, Hassan, R., Scholes, R., and Ash, N., eds Island Press, p. 513-550. +Delille, B., Harlay, J., Zondervan, I., Jacquet, S., Chou, L., et al. (2005). Response o primary production and calcification to changes of pCO2 during experimenta blooms of the coccolithophorid Emiliania huxleyi. Global Biogeochemical Cycles 19, GB2023, doi:10.1029/2004GB002318. +Dennison, W.C., Orth, R.J., Moore, K.A., Stevenson, J.C., Carter, V., Kollar, S. Bergstrom, P.W., and Batiuk, R.A. (1993). Assessing Water Quality wit Submersed Aquatic Vegetation. Habitat requirements as barometers o Chesapeake Bay health. BioScience, 43: 86-94. +© 2016 United Nations 3 + +Depledge, M.H., Galgani, F., Panti, C., Caliani, I., Casini, S., and Fossi, M.C. (2013). Plasti litter in the sea. Marine Environmental Research, 92: 279-281. +Deutsch, C., Sarmiento, J.L., Sigman, D.M., Gruber, N., and Dunne, J.P. (2007). Spatia coupling of nitrogen inputs and losses in the ocean. Nature, 445: 163-167. +DeVries, T., Deutsch, C., Primeau, F., Chang, B., and Devol, A. (2012). Global rates o water-column denitrification derived from nitrogen gas measurements. Natur Geoscience, 5: 547-550. +DeVries, T., Deutsch, C., Rafter, P.A., and Primeau, F. (2013). Marine denitrification rate determined from a global 3-dimensional inverse model. Biogeosciences, 10 2481-2496. +deYoung, B., Heath, M., Werner, F., Chai, F., Megrey, B., and Monfray, P. (2004) Challenges of Modeling Ocean Basin Ecosystems. Science, 304: 1463-1466. +Diaz, R.J., and Rosenberg, R. (2008). Spreading Dead Zones and Consequences fo Marine Ecosystems. Science. 321: 926-929. +Donato, D.C., Kauffman, J.B., Murdiyarso, D., Kurnianto, S., Stidham, M., an Kanninen, M. (2011). Mangroves among the most carbon-rich forests in th tropics. Nature Geoscience, 4: 293-297. +Doney, S.C. (2010). The Growing Human Footprint on Coastal and Open-Ocea Biogeochemistry. Science, 328: 1512-1516. +Doney, S.C., Ruckelshaus, M., Duffy, J.E., Barry, J.P., Chan, F., English, C.A. Galindo, H.M., Grebmeier, J.M., Hollowed, A.B., Knowlton, N., Polovina, J. Rabalais, N.N., Sydeman, W.J., and Talley, L.D. (2012). Climate change impacts o marine ecosystems. Annual Reviews - Marine Science, 4: 11-37. +Doney, S.C., Balch, W.M., Fabry, V.J., and Feely, R.A. (2009). Ocean Acidification: Critical Emerging Oroblem for the Ocean Sciences. Oceanography, 22 (4): 16-25. +Duarte, C.M. (1989). Temporal biomass variability and production/biomass relationship of seagrass communities. Marine Ecology Progress Series, 51: 269-276. +Duarte, C.M. (1995). Submerged aquatic vegetation in relation to different nutrien regimes. Ophelia, 41: 87-112. +Duarte, C.M., and Kirkman, H. (2001). Methods for the measurement of seagras abundance and depth distribution. In Global Seagrass Research Methods, Short F.T., and Coles, R.G., eds. 141-153. Elsevier Science B.V. +Duarte, C.M., Borum, J., Short, F.T., and Walker, D.I. (2008). Seagrass ecosystems: thei global status and prospects, p. 281-306. In Aquatic Ecosystems, Polunin, N., ed Cambridge University Press, Cambridge, U.K. +Duarte, C.M., Middelburg, J.J., and Caraco, N.F. (2005). Major role of marine vegetatio on the oceanic carbon cycle. Biogeosciences, 2: 1-8. +© 2016 United Nations 3 + +Duarte, C.M., Marba, N., Gacia, E., Fourqurean, J.W., Beggins, J., Barron, C., an Apostolaki, E.T. (2010). Seagrass community metabolism: Assessing the carbo sink capacity of seagrass meadows. Global Biogeochemical Cycles, 24 (4) GB4032, doi:10.1029/2010GB003793 +Duce, R.A., LaRoche, J., Altieri, K., Arrigo, K.R., Baker, A.R., et al. (2008). Impacts o Atmospheric Anthropogenic Nitrogen on the Open Ocean. Science, 320: 893-897. +Ducklow, H.W., Fraser, W.R., Meredith, M.P., et al. (2013). West Antarctic Peninsula: A Ice-Dependent Coastal Marine Ecosystem in Transition. Oceanography, 26 (3) 190 — 203. +Duke, N.C., Meynecke, J-O., Dittmann, S., Ellison, A.M., Anger, K., et al. (2007). A worl without mangroves? Science, 317: 41-42. +Edwards, M., Reid, P.C., and Planque, B. (2001). Long-term and regional variability o phytoplankton biomass in the north-east Atlantic (1960-1995). /CES Journal o Marine Science 58: 39-49. +Edwards, M., and Richardson, A.J. (2004). Impact of climate change on marine pelagi phenology and trophic mismatch. Nature, 430: 881-884. +Edwards, M., Johns, D.G., Leterme, S.C., Svendsen, E., and Richardson, A.J. (2006) Regional climate change and harmful algal blooms in the northeast Atlantic Limnology and Oceanography, Inc., 51 (2): 820-829. +Edwards, M., Beaugrand, G., Johns, D.G., Licandro, P., McQuatters-Gollop, A., an Wootton, M. (2010). Ecological status report: results from the CPR survey 2009 SAHFOS Technical Report, 7: 1-8. Plymouth, U.K. ISSN 1744-0750. +Emerson, S., and Hedges, J.1. (2008). Chemical Oceanography and the Marine Carbo Cycle, Cambridge University Press, Cambridge, U.K. +Engel, I., Luo, B.P., Pitts, M.C., Poole, L.R., Hoyle, C.R., Groof, J.-U., Dornbrack, A., an Peter, T. (2013). Heterogeneous formation of polar stratospheric clouds — Part 2 Nucleation of ice on synoptic scales. Atmospheric Chemistry and Physic Discussions, 13: 8831-8872. +Falkowski, P.G. (1997), Evolution of the nitrogen cycle and its influence on the biologica sequestration of CO, in the ocean, Nature, 387: 272-275. +Falkowksi P.G., Katz, M.E., Knoll, A.H., Quigg, A., Raven, J.A., Schofield, O., an Taylor, F.J.R. (2004). The Evolution of Modern Eukaryotic Phytoplankton. Science 305 (5682): 354-360. +Falkowski, P.G., and Raven, J.A. (1997). Aquatic photosynthesis. Blackwell Science Malden, MA. +Falkowski, P.G. Barber, R.T., and Smetacek, V. (1998). Biogeochemical Controls an Feedbacks on Ocean Primary Production. Science, 281: 200-206. +© 2016 United Nations 4 + +Feely, R.A., Doney, S.C., and Cooley, S.R. (2009). Ocean Acidification: Present Condition and Future Changes in a High-CO2 World. Oceanography, 22 (4): 36-47. +Ferreyra, G.A., Mostajir, B., Schloss, I.R., Chatila, K., Ferrario, M.E., Sargian, P., Roy, S. Prod'homme, J., and Demers. S. (2006). Ultraviolet-B radiation effects on th structure and function of lower trophic levels of the marine planktonic food web Photochem Phtobiol, 82, 887-897. +Field, C.B., Behrenfeld, M.J., Randerson, J.T., and Falkowski, P. (1998). Primar Production of the Biosphere: Integrating Terrestrial and Oceanic Components Science, 281, 237-240. +Filippelli, G.M., and Delaney, M.L. (1996). Phosphorus geochemistry of equatorial Pacifi sediments. Geochimica et Cosmochimica Acta, 60:1479-1495. +Finkel, Z.V., Beardall, J., Flynn, K.J., Quigg, A., Rees, T.A.V., and Raven J.A. (2010) Phytoplankton in a changing world: cell size and elemental stoichiometry Journal of Plankton Research, 32: 119-137. +Flores, H., Atkinson, A., Kawaguchi, S., Krafft, B.A., Milinevsky, G., Nicol, S., et al. (2012) Impact of climate change on Antarctic krill. Marine Ecology Progress Series, 458 1-19. +Forget, M-H., Stuart, V., and Platt, T., eds. (2009). Remote Sensing in Fisheries an Aquaculture. [OCCG Report. No. 8, 120 pp. +Forget, M-H., Platt, T., Sathyendranath, S., Stuart, V., and Delaney, L. (2010). "Societa Applications in Fisheries and Aquaculture Using Remotely-Sensed Imagery - Th SAFARI Project" in Proceedings of OceanObs’09: Sustained Ocean Observation and Information for Society (Vol. 2), Venice, Italy, 21-25 September 2009, Hall, J. Harrison, D.E., and Stammer, D., eds., ESA Publication WPP-306 doi:10.5270/OceanObs09.cwp.30. +Fourqurean, J.W., Duarte, C.M., Kennedy, H., Marba, N., Holmer, M., Mateo, M.A. Apostolaki, E.T., Kendrick, G.A., Krause-Jensen, D., McGlathery, K.J., and Serrano O. (2012). Seagrass ecosystems as a globally significant carbon stock. Natur Geoscience, 5: 505-509. +Fraile, I., Schulz, M., Mulitza, S., and Kucera, M. (2008). Predicting the global distributio of planktonic foraminifera using a dynamic ecosystem model. Biogeosciences, 5 891-911. +Frank, K.T., Petrie, B., Shackell, N.L., and Choi, J.S. (2006). Reconciling differences i trophic control in mid-latitude marine ecosystems. Ecology Letters, 9: 1096 1105. +Franks, P.J.S. (2008). Physics and physical modeling of harmful algal blooms. In Babin M., Roesler, C.S., and Cullen, J.J., eds. Real-Time Coastal Observing Systems fo Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory, Instrumentation +© 2016 United Nations 4 + +and Modelling. Monographs on Oceanographic Methodology Series, UNESCO. pp. 561-598. +Freing, A., Wallace, D.W.R., and Bange, H.W. (2012). Global oceanic production o nitrous oxide. Philosophical Transactions of the Royal Society B, 367: 1245-1255. +Fricke, A., Molis, M., Wiencke, C., Valdivia, N., and Chapman, A.S. (2011). Effects of U radiation on the structure of Arctic macrobenthic communities. Polar Biology 34, 995-1009. +Friedland, K.D., Stock, C., Drinkwater, K.F., Link, J.S., Leaf, R.T., et al. (2012). Pathway between Primary Production and Fisheries Yields of Large Marine Ecosystems PLoS ONE, 7(1): 28945 doi: 10.1371/journal.pone.0028945. +Friedrichs M.A.M, Carr, M-E., Barber, R.T., Scardi, M., Antoine, D., et al. (2009) Assessing the uncertainties of model estimates of primary productivity in th tropical Pacific Ocean. Journal of Marine Systems, 76: 113-133. +Galloway, J.N., Dentener, F.J., Capone, D.G., Boyer, E.W., Howarth, R.W., Seitzinger, S.P. Asner, G.P., Cleveland, C.C., Green, P.A., Holland, E.A., Karl, D.M., Michaels, A.F. Porter, J.H., Townsend, A.R., and Vérs6marty, C.J. (2004). Nitrogen cycles: past present and future. Biogeochemistry, 70:153-226. +Garrard S. L., and N. J. Beaumont. (2014). The effect of ocean acidification on carbo storage and sequestration in seagrass beds; a global and UK context. Marin Pollution Bulletin, 86: 138-146. +Geider, R.J., Delucia, E.H., Falkowski, P.G., Finzi, A.C., Grime, J.P., Grace, J., Kana, T.M. La Roche, J., Long, S.P., Osborne, B.A., Platt, T., Prentice, I.C., Raven, J.A. Schlesinger, W.H., Smetacek, V., Stuart, V., Sathyendrananth, S., Thomas, R.B. Vogelmann, T.C., Williams, P., and Woodward, F.I. (2001). Primary productivity o planet earth: biological determinants and physical constraints in terrestrial an aquatic habitats. Global Change Biology, 7: 849-882. +Gilman, E.L., Ellison, J., Duke, N.C., and Field, C. (2008). Threats to mangroves fro climate change and adaptation options: a review. Aquatic Botany, 89: 237-250. +Gitelson, A.A. (2004). Wide dynamic rage vegetation index for remote quantification o biophysical characteristics of vegetation. Journal of plant physiology, 161: 165-173. +Glibert, P.M., Anderson, D.M., Gentien, P., Granéli, E., and Sellner, K.G. (2005). Th global, complex phenomena of harmful algal blooms. Oceanography, 18(2): 136 147. +Glibert, P.M., and Bouwman, L. (2012). Land-based Nutrient Pollution and th Relationship to Harmful Algal Blooms in Coastal Marine Systems. Inprint, 2: 5- (www.loicz.org). +Goes, J.I., Handa, N., Taguchi, S., and Hama, T. (1995). Changes in the patterns o biosynthesis and composition of amino acids in a marine phytoplankton exposed to +© 2016 United Nations 4 + +ultraviolet-B radiation: nitrogen limitation implicated. Photochemistry an Photobiology, 62, 703-710. +Graham, W.M., Gelcich, S., Robinson, K.L., Duarte, C.M., Brotz, L., Purcell, J.E. Madin, L.P., Mianzan, H., Sutherland, K.R., Uye, S.-I., and others. (2014). Linkin human well-being and jellyfish: Ecosystem services, impacts and societal responses Frontiers in the Ecology and Environment, 12: 515-523. +Gregg, W.W., Conkright, M.E., Ginoux, P., O’Reilly, J.E., and Casey, N.W. (2003). Ocea primary production and climate: global decadal changes. Geophysical Researc Letters, 30 (15): 1809, doi:10.1029/2003GL016889. +Gregg, W.W., Casey, N.W., O'Reilly, J.E., and Esaias, W.E. (2009). An empirical approac to ocean color data: Reducing bias and the need for post-launch radiometric re calibration. Remote Sensing of Environment, 113: 1598-1612. +Greene, C.H., and Pershing, A.J. (2004). Climate and the Conservation Biology of Nort Atlantic Right Whales: The Right Whale at the Wrong Time? Frontiers in Ecolog and the Environment, 2: 29-34. +Gross, M.F., Klemas, V., and Hardisky, M.A. (1990). Long-term remote monitoring of sal marsh biomass. Proceedings SPIE 1300, Remote Sensing of the Biosphere, 59 doi:10.1117/12.21390. +GroBkopf, T., Mohr, W., Baustian, T., Schunck, H., Gill, D., Kuypers, M.M.M., Lavik, G. Schmitz, R.A., Wallace, D.W.R., and LaRoche, J. (2012). Doubling of marin dinitrogen-fixation rates based on direct measurements. Nature, 488, 361-364. +Gruber, N. (2004). The Dynamics of the Marine Nitrogen Cycle and its Influence o Atmospheric CO) Variations. In The Ocean Carbon Cycle and Climate, NAT Science Series, Follows, M., and Oguz. T., eds. Dordrecht: Kluwer Academic, pp 97-148. +Gruber, N. (2008). The Marine Nitrogen Cycle: Overview and Challenges. In Nitrogen i the marine environment. Capone, D.G., Bronk, D.A., Mulholland, M.R., an Carpenter, E.J., eds. Chapter 1, 2"! Edition, San Diego, CA, Academic Press. +Gruber, N., and Sarmiento, J.L. (2002). Large-Scale Biogeochemical—Physical Interaction in Elemental Cycles. In The Sea: Biological-Physical Interactions. Robinson, A.R. McCarthy, J.F., Rothschild, B., eds. (Wiley, New York), vol. 12, pp. 337-399. +Gruber, N., and Galloway, J.N. (2008). An Earth-system perspective of the globa nitrogen cycle. Nature, 451: 293 -296. +Gruber, N., C. Hauri, and G-K. Plattner. (2009). High vulnerability of eastern boundar upwelling systems to ocean acidification. Global Change Newsletter (IGBP), Issu No. 73. +Guidi, L., Degl’Innocenti, E., Remorini, D., Biricolti, S., Fini, A., Ferrini, F., Nicese, F.P., an Tattini, M. (2011). The impact of UV-radiation on the physiology and +© 2016 United Nations 4 + +biochemistry of Ligustrum vulgare exposed to different visible-light irradiance Environmental and Experimental Botany, 70, 88-95. +Ha, S-Y., Joo, H-M., Kang, S-H., Ahn, I-Y., and Shin, K-H. (2014). Effect of ultraviole irradiation on the production and composition of fatty acids in plankton in a sub Antarctic environment. Journal of Oceanography, 70: 1-10. DOI 10.1007/s10872 013-0207-3. +Ha, S-Y., Kim, Y-N., Park, M-O., Kang, S-H., Kim, H-C. and Shin, K-H. (2012). Production o mycosporine-like amino acids of in situ phytoplankton community i Kongsfjorden, Svalbard, Arctic. Journal of Photochemistry and Photobiology, B Biology, 114, 1-14. +Haddad, N.M., Crutsinger, G.M., Gross, K., Haarstad, J., and Tilman, D. (2011). Plan diversity and the stability of foodwebs. Ecology Letters, 14: 42-46. +Hader, D.-P., Kumar, H.D., Smith, R.C., and Worrest, R.C. (2007). Effects of solar U radiation on aquatic ecosystems and interactions with climate change Photochemical and Photobiological Sciences, 6, 267-285. +Hall, J., Harrison, D.E., and Stammer, D., eds. (2010). Proceedings of OceanObs’09 Sustained Ocean Observations and Information for Society (Volumes 1 and 2) ESA Publication WPP-306, doi:10.5270/OceanObs09. +Hallegraeff, G.M. (2010). Ocean climate change, phytoplankton community responses and harmful algal blooms: a formidable predictive challenge. Journal of th Phycology, 46: 220-235. +Hamersley, M.R., Lavik, G., Woebken, D., Rattray, J.E., Lam, P., Hopmans, E.C. Damsté, J.S.S., Kriiger, S., Graco, M., Gutiérrez, D., and Kuypers, M.M.M. (2007) Anaerobic ammonium oxidation in the Peruvian oxygen minimum zone Limnology and Oceanography, 52 (3): 923-933. +Xu, H., Miao, X., and Wu, Q. (2006). High quality biodiesel production from a macroalg Chlorella protothecoides by heterotrophic growth in fermenters. Journal o Biotechnology 126:499-507. +Handy, R.D., Owen, R., and Valsami-Jones, E. (2008). The ecotoxicology of nanoparticle and nanomaterials: current status, knowledge gaps, challenges, and futur needs. Ecotoxicolog 17:315-325. +Harrison, J.A., Bouwman, A.F., Mayorga, E., and Seitzinger, S. (2010). Magnitudes an sources of dissolved inorganic phosphorus inputs to surface fresh waters and th coastal zone: a new global model. Global Biogeochemical Cycles, 24, GB1003 doi: 10.1029/2009GB003590. +Hassan, R., Scholes, R., and Ash, N., eds. (2005). Coastal ecosystems In Ecosystems an Human Well-being: Current State and Trends, Island Press. +© 2016 United Nations 4 + +Hay, S. (2006). Marine ecology: gelatinous bells may ring change in marine ecosystems Current Biology 16(17): R679-82. +Hays G.C., Richardson, A.J., and Robinson, C. (2005). Climate change and marin plankton. Trends in Ecology & Evolution, 20 (6): 337-344. +Hay, S. (2006). Marine Ecology: Gelatinous Bells May Ring Change in Marine Ecosystems Current Biology, 16: 679-682. +Heck, Jr., K.L., Carruthers, T.J.B., Duarte, C.M., Hughes, A.R., Kendrick, G., Orth, R.J., an Williams, S.W. (2008). Trophic Transfers from Seagrass Meadows Subsidiz Diverse Marine and Terrestrial Consumers. Ecosystems, 11: 1198-1210. +Henley, S.F. (2013). Climate-induced changes in carbon and nitrogen cycling in th rapidly warming Antarctic coastal ocean. (http://hdl.handle.net/1842/7626). +Henson, S.A., and Thomas, A.C. (2007). Phytoplankton Scales of Variability in th California Current System: 1. Interannual and Cross-Shelf Variability, Journal o Geophysical Research-Oceans, 112, CO7017, doi:10.1029/2006JC004039. +Henson, S.A., Sarmiento, J.L., Dunne, J.P., Bopp, L., Lima, I., Doney, S.C., John, J., an Beaulieu, C. (2010). Detection of anthropogenic climate change in satellit records of ocean chlorophyll and productivity. Biogeosciences, 7: 621-640. +Henson, S.,Lampitt, R., and Johns, D. (2012). Variability in phytoplankton communit structure in response to the North Atlantic Oscillation and implications fo organic carbon flux. Limnology and Oceanography, 57(6): 1591-1601. +Henson, S., Cole, H., Beaulieu, C., and Yool, A. (2013). The impact of global warming o seasonality of ocean primary production. Biogeosciences, 10: 4357-4369 +Heumann, B.W. (2011). Satellite remote sensing of mangrove forests: Recent advance and future opportunities. Progress in Physical Geography, 35 (1): 87-108. +Hilligsge, K.M., Richardson, K., Bendtsen, J., S#rensen, L-L., Nielsen, T.G., an Lyngsgaard, M.M. (2011). Linking phytoplankton community size compositio with temperature, plankton food web structure and sea-air CO2 flux. Deep-Se Research Part I: Oceanographic Research Papers, +58: 826-838. +Hoagland P., and Scatasta, S. (2006). The Economic Effects of Harmful Algal Blooms. I Granéli, E., and Turner, J.T., eds. Ecology of Harmful Algae. Ecology Studie Series: Springer-Verlag Berlin Heidelber. 189: 391-402. +Hofmann, L.C., Bischof, K. (2014). Ocean acidification effects on calcifying macroalgae Aquatic Biology, 22: 261-279. +Hood, M., ed. (2009). Ship-based Repeat Hydrography: A Strategy for a Sustained Globa Program. UNESCO-IOC Technical Series, 89. IOCCP Reports, 17. ICPO Publicatio 142. (www.go-ship.org/Docs/IOCTS89_GOSHIP.pdf). +© 2016 United Nations 4 + +Hooper, D.U., Chapin Ill, F.S., Ewel, J.J., Hector, A., Inchausti, P., Lavorel, S., Lawton, J.H. Lodge, D.M., Loreau, M., Naeem, S., Schmid, B., Setala, H., Symstad, A.J. Vandermeer, J., and Wardle, D.A. (2005). Effects of biodiversity on ecosyste functioning: A consensus of current knowledge. Ecological Monographs, 75: 3 35. +Howarth, R.W., et al. (1996). Regional nitrogen budgets and riverine N and P fluxes fo the drainages to the North Atlantic Ocean: Natural and human influences Biogeochemistry, 35, 75— 139. +Howarth, R.W. (2008). Coastal nitrogen pollution: A review of sources and trend globally and regionally. Harmful Algae, 8: 14-20. +Howarth, R.W., Chan, F., Conley, D.J., Garnier, J., Doney, S.C., Marino, R., and Billen, G (2011). Coupled biogeochemical cycles: eutrophication and hypoxia in temperat estuaries and coastal marine ecosystems. Frontiers in Ecology and th Environment, 9(1): 18-26. +Howarth, R.W., Swaney, D.P., Billen, G., Garnier, J., Hong, B., Humborg, C., Johnes, P.J. Morth, C-M., and Marino, R. (2012). Nitrogen fluxes from large watersheds t coastal ecosystems controlled by net anthropogenic nitrogen inputs and climate Frontiers in Ecology and the Environment. 10(1): 37-43. +Hu, C., Feng, L., Lee, Z., Davis, C.0., Mannino, A., McClain, C.R., and Franz, B.A. (2012) Dynamic range and sensitivity requirements of satellite ocean color sensors learning from the past. Applied Optics, 51 (25): 6045-6062. +Huete-Ortega, M., Marafion, E., Varela, M., Bode, A. (2010). General patterns in the siz scaling of phytoplankton abundance in coastal waters during a 10-year tim series. Journal of Plankton Research, 32: 1-14. +Hutchins, D.A., Mulholland, M.R., and Fu, F-X. (2009). Nutrient Cycles and Marin Microbes in CO2-Enriched Ocean. Oceanography, 22: 128-145. +IOCCG. (1998). Minimum Requirements for an Operational Ocean-Colour Sensor for th Open Ocean. Report 1, Morel, A., ed. International Ocean-Colour Coordinatin Group (IOCCG) Report 1, pp.48. +lrigoien, X., Huisman, J., and Harris, R.P. (2004). Global biodiversity patterns of marin phytoplankton and zooplankton. Nature, 429: 863-867. +Jang, C.J., Park, J., Park, T., and Yoo, S. (2011) Response of the ocean mixed layer dept to global warming and its impact on primary production: a case for the Nort Pacific Ocean. ICES Journal of Marine Science. doi:10.1093/icesjms/fsr064. +Jansen, T., Campbell, A., Brunel, T., Clausen, L.M.. (2013). Spatial Segregation within th Spawning Migration of North Eastern Atlantic Mackerel (Scomber scombrus) a Indicated by Juvenile Growth Patterns. PloS One, doi 10.1371/journal.pone.0058114. +© 2016 United Nations 4 + +Jia, M., Wang, Z., Liu, D., Ren, C., Tang, X., and Dong, Z. (2013). Monitoring Loss an Recovery of Salt Marshes in the Liao River Delta, China. Journal of Coasta Research (in-press). 31: 371-377. +Doi: http://dx.doi.org/10.2112/JCOASTRES-D-13-00056.1. +Jiang, H., and Qiu, B. (2011). Inhibition of photosynthesis by UV-B exposure and it repair in the bloom-forming cyanobacterium Microcystis aeruginosa. Journal o Applied Phycology, 23 (4): 691-696. +Johnson, K.S., and Coletti, L.J. (2002). /n situ ultraviolet spectrophotometry for hig resolution and long-term monitoring of nitrate, bromide and bisulfide in th ocean. Deep-Sea Research Part |: Oceanography Research Papers, 49: 1291-1305. +Kaldy, J.E., and Dunton, K.H. (2000). Above- and below-ground production, biomass an reproductive ecology of Thalassia testudinum (turtle grass) in a subtropica coastal lagoon. Marine Ecology Progress Series, 193: 271-283. +Kamykowski, D. (1987). A preliminary biophysical model of the relationship betwee temperature and plant nutrients in the upper ocean. Deep Sea Research Part A Oceanographic Research Papers, +34 (7): 1067-1079. +Karl, D.M., Christian, J.R., Dore, J.E., Hebel, D.V., Letelier, R.M., Tupas, L.M., an Winn, C.D. (1996). Seasonal and interannual variability in primary production an particle flux at Station ALOHA Deep Sea Research Part II: Topical Studies in Oceanography, 43: 539-568. +Karl, D., Michaels, A., Bergman, B., Capone, D., Carpenter, E., Letelier, R., Lipschultz, F. Paerl, H., Sigman, D., and Stal, L. (2002). Dinitrogen fixation in the world’ oceans. Biogeochemistry, +57/58: 47-98. +Keeling, R.F., Kortzinger, A., and Gruber, N. (2010). Ocean deoxygenation in a warmin world. Annual Reviews Marine Science, 2: 199-229. +Kemp, W.M., Lewis, M.R., Jones, T.W. (1986). Comparison of methods for measurin productlon by the submersed macrophyte, Potamogeton perfoliatus. Limnolog and Oceanography, 31: 1322-1334. +Kemp W.M., Testa, J.M., Conley, D.J., Gilbert, D., and Hagy, J.D. (2009). Tempora responses of coastal hypoxia to nutrient loading and physical controls Biogeosciences, 6: 2985-3008. +Kemp, W.M., Batiuk, R., Bartleson, R., Bergstrom, P., Carter, V., Gallegos, G., Hunley, W. Karrh, L., Koch, E., Landwehr, J., Moore, K., Murray, L., Naylor, M., Rybicki, N. Stevenson, J.C., and Wilcox, D. (2004). Habitat Requirements for Submerge Aquatic Vegetation in Chesapeake Bay: Water Quality, Light Regime, an Physical-Chemical Factors. Estuaries, 27: 363-377. +Kideys, A. E. (2002). Fall and Rise of the Black Sea Ecosystem. Science, 297: 1482-1484. +© 2016 United Nations 4 + +Kigrboe T. (1993). Turbulence, phytoplankton cell size, and the structure of pelagic foo webs. Advances in Marine Biology 29:72. +Koch, E.W., Barbier, E.B., Silliman, B.R., Reed, D.J., Perillo, G.M.E., Hacker, S.D. Granek, E.F., Primavera, J.H., Muthiga, N., Polasky, S., Halpern, B.S. Kennedy, C.J., Kappel, C.V., and Wolanski, E. (2009). Non-linearity in ecosyste services: temporal and spatial variability in coastal protection. Frontiers i Ecology and the Environment, 7(1): 29-37, doi:10.1890/080126. +Koeller, P., Fuentes-Yaco, C., Platt, T., Sathyendranath, S., Richards, A., Ouellet, P. Orr, D., Skuladottir, U., Wieland, K., Savard, L., and Aschan, M. (2009). Basin Scale Coherence in Phenology of Shrimps and Phytoplankton in the Nort Atlantic Ocean. Science, 324:791-793. +Koeve, W., and Kahler, P.K. (2010). Heterotrophic denitrification vs. autotrophi anammox — quantifying collateral effects on the oceanic carbon cycle Biogeosciences, 7: 2327-2337. +Kovacs, J.M., King, J.M.L., de Santiago, F.F., Flores-Verdugo. F. (2009). Evaluating th condition of a mangrove forest of the Mexican Pacific based on an estimated lea area index mapping approach. Environmental Monitoring and Assessment, 157 137-149. +Kovacs, J., Wang, M.J., Blanco-Correa, M.. (2001). Mapping mangrove disturbance using multi-date Landsat TM imagery. Environmental Management, 27: 763-776. +Krishnamurthy, A., Moore, J. K., Mahowald, N., Luo, C., Doney, S.C., Lindsay, K. Zender, C.S. (2009). Impacts of increasing anthropogenic soluble iron an nitrogen deposition on ocean biogeochemistry. Global Biogeochemical Cycles 23, GB3016, doi:10.1029/2008GB003440. +Krishnamurthy, A., Moore, J.K., Mahowald, N., Luo, C., and Zender. C.S. (2010). Impact of atmospheric nutrient inputs on marine biogeochemistry. Journal o Geophysical Research, 115: 13 pp., GO1006, doi:10.1029/2009JG001115. +Kroeker, K.J., Kordas, R.L., Crim, R.N., and Singh, G.G. (2010). Meta-analysis reveal negative yet variable effects of ocean acidification on marine organisms. Ecolog Letters, 13: 1419-1434. +Kroeze, C., and Seitzinger, S.P. (1998). Nitrogen inputs to rivers, estuaries an continental shelves and related nitrous oxide emissions in 1990 and 2050: global model. Nutrient Cycling in Agroecosystems, 52: 195-212. +Laffoley, D., and Grimsditch, G. (2009). The management of natural coastal carbon sinks Gland: IUCN. +Lam, P., Lavik, G., Jensen, M.M., van de Vossenberg, J., Schmid, M., Woebken, D. Gutiérrez, D., Amann, R., Jetten, M.S.M., and Kuypers, M.M.M. (2009). Revisin the nitrogen cycle in the Peruvian oxygen minimum zone. Proceedings of the +© 2016 United Nations 4 + +National Academy of Sciences (USA) 106: 4752-4757. +Lamarque, J.-F., Dentener, F., McConnell, J., et al. (2013). Multi-model mean nitroge and sulfur deposition from the Atmospheric Chemistry and Climate Mode Intercomparison rojec (ACCMIP): evaluation of historical and projected future changes. Atmospheri Chemistry and Physics, 13: 7997-8018. +Landolfi, A., Dietze, H., Koeve, W., and Oschlies, A. (2013). Overlooked runawa feedback in the marine nitrogen cycle: the vicious cycle. Biogeosciences, 10 1351-1363. +Landry, M.R., Barber, R.T., Bidigare, R.R., Chai, F., Coale, K.H., et al. (1997). lron an grazing constraints on primary production in the central equatorial Pacific: An EqPa synthesis. Limnology and Oceanography, 42: 405-418. +Laufk6étter, C., Vogt, M., and Gruber, N. (2013). Long-term trends in ocean plankto production and particle export between 1960-2006. Biogeosciences, 10: 7373 7393. +Laws, E.A., Falkowski, P.G., Smith, Jr., W.O., Ducklow, H., and McCarthy, J.J. (2000) Temperature effects on export production in the open ocean. Globa Biogeochemical Cycles, 14: 1231-1246. +Lebrato, M., and Jones, D.O.B. (2011). Jellyfish biomass in the biological pump Expanding the oceanic carbon cycle. The Biochemical Society Journal, p. 35-39. +Lee, Z.-P., ed. (2006). Remote sensing of inherent optical properties: Fundamentals Tests of algorithms, and applications. [OCCG Report No. 5, 126 pp., Dartmouth Canada. +Legendre, L., and Rassoulzadegan, F. (1996). Food-web mediated export of biogeni carbon in oceans: hydrodinamic control. Marine Ecology Progress Series, 145 179-193. +Le Quéré, C., Takahashi, T., Buitenhuis, E.T., Rodenbeck, C., and Sutherland, S.C. (2010) Impact of climate change and variability on the global oceanic sink of CO2. Globa Biogeochemical Cycles, +24 (4): GB4007, doi: 10.1029/2009GB003599. +Licandro, P., Conway, D.V.P., Daly Yahia, M.N., Fernandez de Puelles, M.L., Gasparini, S. Hecq, J.H., Tranter, P., Kirby, R.R. (2010). A blooming jellyfish in the northeas Atlantic and Mediterranean. Biology Letters, doi:10.1098/rsbl.2010.0150. +Lilly, E.L., Halanych, K.M., and Anderson, D.M. (2007). Species boundaries and globa biogeography of the Alexandrium tamarense complex (Dinophyceae). Journal o Phycology, 43: 1329-1338. +© 2016 United Nations 4 + +Liu, J., Liu, S., Loveland, T.R., Tieszen, L.L. (2008). Integrating remotely sensed land cove observations and a biogeochemical model for estimating forest ecosyste carbon dynamics. Ecological Modelling, 219: 361-372. +Lohrenz, S.E., Fahnensteil, G.L., Redalje, D.G., Lang, G.A., Chen, X., and Dagg, M.J. (1997) Variations in primary production of northern Gulf of Mexico continental shel waters linked to nutrient inputs from the Mississippi River. Marine Ecolog Progress Series, 155: 45-54. +Lomas, M.W., Bates, N.R., Johnson, R.J., Knap, A.H., Steinberg, D.K., and Carlson, C.A (2013). Two decades and counting: 24-years of sustained open ocea biogeochemical measurements in the Sargasso Sea. Deep Sea Research Part II Topical Studies in Oceanography. 93: 16-32 doi: 10.1016/j.dsr2.2013.01.008. +Long, S.P, Jones, M.B., and Roberts, M.J., eds. (1992). Primary Productivity of Gras Ecosystems of the Tropics and Sub-tropics. Chapman and Hall, London, 267 pp. +Luo, Y.-W., Doney, S.C., Anderson, L.A., Benavides, M., Bode, A., et al. (2012). Databas of diazotrophs in global ocean: abundances, biomass and nitrogen fixation rates Earth System Science Data, 4: 47-73. +Lynam C. P., Gibbons, M.J., Axelsen, B.E., Sparks, C.A.J., Coetzee, J., Heywood, B.G. Brierley, A.S. (2006). Jellyfish overtake fish in a heavily fished ecosystem. Curr Biol., 16: 492-493. +Ma, S., Tao, Z., Tang, X., Yu, Y., Zhou, X., Ma, W., and Li, Z. (2014). Estimation of Marin Primary Productivity From Satellite-Derived Phytoplankton Absorption Data. /EE Journal of Selected Topics in Applied Earth Observations and Remote Sensing, (7): 3084-3092. +Mackas, D.L. (2011). Does blending of chlorophyll data bias temporal trend? Nature 472: E4-E5. +Mackas, D.L., and Beaugrand, G. (2010). Comparisons of zooplankton time series Journal of Marine Systems, 79: 286-304. +Mackas, D.L., Batten, S.D., and Trudel, M. (2007). Effects on zooplankton of a warmin ocean: recent evidence from the Northeast Pacific. Progress in Oceanography 75(2): 223-252. +MacKenzie, R.A., and Dionne, M. (2008). Habitat heterogeneity: importance of sal marsh pools and high marsh surfaces to fish production in two Gulf of Maine sal marshes. Marine Ecology Progress Series, 368: 217-230. +Mahaffey, C., Michaels A.F., Capone, D.G. (2005). The conundrum of marine N> fixation American Journal of Science, 305 (6-8): 546-595. doi: 10.2475/ajs.305.6-8.546. +Mahowald, N.M., Baker, A.R., Bergametti, G., Brooks, N., Duce, R.A., Jickells, T.D. Kubilay, N., Prospero, J.M., and Tegen, I. (2005). Atmospheric global dust cycle +© 2016 United Nations 5 + +and iron inputs to the ocean, Global Biogeochemical Cycles, 19, GB4025 doi:10.1029/2004GB002402. +Mahowald, N., Jickells, T.D., Baker, A., Artaxo, P., Benitez-Nelson, C.R., et al. (2008) Global distribution of atmospheric phosphorus sources, concentrations an deposition rates, and anthropogenic impacts, Global Biogeochemical Cycles, 22 GB4026, doi:10.1029/2008GB003240. +Malone, T.C. (1971). Diurnal rhythms in netplankton and nanoplankton assimilatio ratios. Marine Biology, 10: 285-289. +Malone, T.C. (1980). Algal size and phytoplankton ecology. pp. 433-464. In: Morris, I. ed. The Physiological Ecology of Phytoplankton, Blackwell Scientific Publications London. +Malone, T.C. (1991). River flow, phytoplankton production and oxygen depletion i Chesapeake Bay, In Modern and Ancient Continental Shelf Anoxia, Tyson, R.V. and Pearson, T.H., eds. Geological Society Publication No. 58, p. 83-94. +Malone, T.C. (2008). Ecosystem dynamics, harmful algal blooms and operationa oceanography. In Babin, M., Roesler, C.S., and Cullen, J.J., eds. Real-Time Coasta Observing Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms Monographs on Oceanographic Methodology Series, UNESCO, pp. 527-560. +Malone, T.C., Conley, D.J., Fisher, T.R., Glibert, P.M., Harding, L.W., Sellner, K. (1996) Scales of Nutrient-Limited Phytoplankton Productivity in Chesapeake Bay Estuaries, 19 (2B): 371-385. +Malone, T.C., DiGiacomo, P.M., Goncalves, E., Knap, A.H., Talaue-McManus, L., an de Mora, S. (2014a). A Global Ocean Observing System Framework fo Sustainable Development. Marine Policy, 43:262-272. +Malone, T.C., DiGiacomo, P.M., Goncalves, E.J., Knap, A.H., Talaue-McManus, L. de Mora, S., and Muelbert. J. (2014b). Enhancing the Global Ocean Observin System to meet evidence based needs for ecosystem-based management o coastal ecosystem services. Natural Resources Forum 38: 168-181. +Malone, T.C., Hopkins, T.S., Falkowski, P.G., and Whitledge, T.E. (1983). Production an transport of phytoplankton biomass over the continental shelf of the New Yor Bight. Continental Shelf Research, 1: 305-337. +Malone, T.C., Malej, A., Harding, L.W., Smodlaka, N., eds. (1999). Ecosystems at th Land-Sea Margin: Drainage Basin to the Coastal Sea. AGU, Coastal and Estuarin Studies, No. 55: 377 p. +Mann, K.H. (1972). Ecological energetics of the sea-weed zone in a marine bay on th Atlantic coast of Canada II. Productivity of the seaweeds. Marine Biology, 14 199-209. +© 2016 United Nations 5 + +Marafién, E. (2009). Phytoplankton Size Structure. In Encyclopedia of Ocean Science, 2" Edition, Steele, J.H., Turekian, K.K., and Thorpe, S.A., eds. Academic Press. +Marafion, E., Behrenfeld, M.J., Gonzalez, N., Mourifio, B., and Zubkov, M.V. (2003). Hig variability of primary production in oligotrophic waters of the Atlantic Ocean uncoupling from phytoplankton biomass and size structure. Marine Ecolog Progress Series, 257: 1-11 +Margalef, R. (1978). Life-forms of phytoplankton as survival alternatives in an unstabl environment. Oceanologica Acta, 1: 493-509. +Marra, J., and Ducklow, H.W. (1995). Primary Production in the North Atlantic Measurements, Scaling, and Optical Determinants. Philosophical Transactions Biological Sciences, Royal Society London B, 348: 153-160. +Marra, J. (2002). Approaches to the Measurement of Plankton Production. pp. 78-108 In Williams, P.J.L., Thomas, D.N., and Reynolds, C.S., eds. Phytoplankto Productivity: Carbon Assimilation in Marine and Freshwater Ecosystems Blackwell Science, Oxford. +Martin, J.H., Coale, K.H., Johnson, K.S., Fitzwater, S.E., et al. (1994). Testing the iro hypothesis in ecosystems of the equatorial Pacific Ocean. Nature, 371: 123-129. +Martinez, M.L., Intralawan, A., Vazquez, G., Pérez-Maqueo, O., Sutton, P., an Landgrave, R. (2007). The coasts of our world: Ecological, economic and socia importance. Ecological Economic 63: 254-272. +Martiny, A.C., Pham, C.T.A., Primeau, F.W., Vrugt, J.A., Moore, J.K., Levin, S.A. Lomas, M.W. (2013), Strong latitudinal patterns in the elemental ratios of marin plankton and organic matter. Nature geosciences, 6: 279-283, doi 10.1038/ngeo1757. +Mathis, J.T., and Feely, R.A. (2013). Building an integrated coastal ocean acidificatio monitoring network in the U.S. Elementa: Science of the Anthropocene, 1 00000 doi: 10.12952/journal.elementa.000007. +McCann, K.S. (2000). The diversity—stability debate. Nature, 405: 228-233. +McClain, C.R. (2009). A Decade of Satellite Ocean Color Observations. Annual Review Marine Science, 1: 19-42. +McGregor, H.V., Dima, M., Fischer, H.W., and Mulitza, S. (2007). Rapid 20th-centur increase in coastal upwelling off northwest Africa. Science, 315: 637-639. +McLeod, K., and Leslie, H., eds. (2009). Ecosystem-Based Management for the Oceans Washington, DC: Island Press. 392 pp. +McLeod, E., Chmura, G.L., Bouillon, S., Salm, R., Bjork, M., Duarte, C.M., Lovelock, C.E. Schlesinger, W.H., and Silliman, B.R. (2011). A blueprint for blue carbon: toward +© 2016 United Nations 5 + +an improved understanding of the role of vegetated coastal habitats i sequestering CO>. Frontiers in Ecology and the Environment 9: 552-560. +McQuatters-Gollop, A., Reid, P.C., Edwards, M., Burkill, P.H., Castellani, C., Batten, S. Gieskes, W., et al. (2011). Is there a decline in marine phytoplankton? Nature 472: E6-E7. +Meier, M.F., Dyurgerov, M.B., Rick, U.K., O'Neel, S., Pfeffer, W.T., Anderson, R.S. Anderson, S.P., and Glazovsky, A.F. (2007). Glaciers Dominate Eustatic Sea-Leve Rise in the 21* Century. Science 317 (5841): 1064-1067. +Meybeck, M. (1982). Carbon, Nitrogen, and Phosphorus Transport by World Rivers American Journal of Science, 282: 401-450. +Millennium Ecosystem Assessment. (2005). Ecosystems and Human Well-being Synthesis. Island Press. +Miller, R.J., Bennett, S., Keller, A.A., Pease, S., and Lenihan, H.S. (2012). TiO Nanoparticles are phototoxic to marine phytoplankton. PLoS ONE 7 (1): e30321. +doi:10.1371/journal.pone.0030321. +Mills, M.M., and Arrigo, K.R. (2010). Magnitude of oceanic nitrogen fixation influence by the nutrient uptake ratio of phytoplankton. Nature Geoscience, 3: 412-416. +Mishra, D.R., Cho, H.J., Ghosh, S., Fox, A., Downs, C., Merani, P.B.T., Kirui, P., et al (2012). Post-spill state of the marsh: Remote estimation of the ecological impac of the Gulf of Mexico oil spill on Louisiana Salt Marshes. Remote Sensing o Environment, 118: 176-185. +Mitsch, W.J., and Gosselink, J.G. (2008). Wetlands. Van Nostrand Reinhold, New York New York, USA. +Moloney, C.L., St John, M.A., Denman, K.L., Karl, D.M., Koster, F.W., Sundby, S., an Wilson, R.P. (2011). Weaving marine food webs from end to end under globa change. Journal of Marine Systems, 84: 106-116. +Montoya, J.M., and Raffaelli, D. (2010). Climate change, biotic interactions an ecosystem services. Philosophical Transactions of the Royal Society B, 365, doi 10.1098/rstb.2010.0114. +Moore, C.M., Mills, M.M., Arrigo, K.R., Berman-Frank, |., Bopp, L., Boyd, P.W. Galbraith, E.D., Geider, R.J., Guieu, C., Jaccard, S.L., Jickells, T.D., La Roche, J. Lenton, T.M., Mahowald, N.M., Marafion, E., Marinov, |., Moore, J.K. Nakatsuka, T., Oschlies, A., Saito, M.A., Thingstad, T.F., Tsuda A., Ulloa O. (2013) Processes and patterns of oceanic nutrient limitation. Nature Geoscience, 6: 701 710, doi: 10.1038/ngeo/1765. +Moore, J.K., Doney, S.C., Lindsay, K., Mahowald, N., Michaels A.F. (2006). Nitroge fixation amplifies the ocean biogeochemical response to decadal timescale +© 2016 United Nations 5 + +variations in mineral dust deposition. Tellus, Series B. 58: 560-572 doi:10.1111/j.1600-0889.2006.00209.x. +Moore, S.K., Trainer, V.L., Mantua, N.J., Parker, M.S., Laws, E.A., Backer, L.C., an Fleming, L.E. (2008). Impacts of climate variability and future climate change o harmful algal blooms and human health. Environmental Health, 7(Suppl 2): $4. +Moran, X.A.G., Lopez-Urrutia, A., Calvo-Diaz, A., and Li, W.K. (2010). Increasin importance of small phytoplankton in a warmer ocean. Global Change Biology 16: 1137-1144. +Morel, A., and Berthon, J-F. (1989). Surface pigments, algal biomass profiles, an potential production of the euphotic layer: Relationships reinvestigated in vie of remote-sensing applications. Limnology and Oceanography, 34: 1545-1562. +Morris, J. T. (2007). Ecological engineering in intertidal saltmarshes. Hydrobiologia, 577 161-168. +Moy, A.D., Howard, W.R., Bray, S.G., and Trull, T.W. (2009). Reduced calcification i modern Southern Ocean planktonic foraminifera. Nature Geoscience, 2: 276-280. +Muller-Karger, F.E., Varela, R., Thunell, R.C., Luerssen, R., Hu, C., Walsh, J.J. (2005). Th importance of continental margins in the global carbon cycle. Geophysica Research Letters, 32, LO1602, doi:10.1029/2004GL021346. +Mustapha, S.B., Bélanger, S., and Larouche, P. (2012). Evaluation of ocean colo algorithms in the southeastern Beaufort Sea, Canadian Arctic: Ne parameterization using SeaWiFS, MODIS, and MERIS spectral bands. Canadia Journal of Remote Sensing, 38(05): 535-556. +Naqvi, W. (2012). Marine nitrogen cycle. The Encyclopedia of Earth (http://www.eoearth.org/view/article/154479). +Naqvi, S.W.A., Voss, M., and Montoya, J.P. (2008). Recent advances in th biogeochemistry of nitrogen in the ocean. Biogeosciences, 5: 1033-1041. +National Research Council. (2004). Groundwater Fluxes across Interfaces. Washington D.C.: National Academies Press. +Negreanu-Pirjol, B., Negreanu-Pirjol, T., Paraschiv, G., Bratu, M., Sirbu, R., Roncea, F. and Meghea, A. (2011). Physical-chemical characterization of some green an red macrophyte algae from the Romanian Black Sea littoral. St. Cerc. St. CICBIA 12 (2): 173-184. +Nellemann, C., Corcoran, E., Duarte, C.M., Valdés, L., De Young, C., Fonseca, L. Grimsditch, G., eds. (2009). Blue Carbon. A Rapid Response Assessment. Unite Nations Environment Programme, GRID-Arendal, www.grida.no. +Nicholls, R.J., and Cazenave, A. (2010). Sea-Level Rise and its Impact on Coastal Zones Science 328 (5985): 1517-1520. +© 2016 United Nations 5 + +Nixon, S.W. (1992). Quantifying the relationship between nitrogen input and th productivity of marine ecosystems. In Takahashi, M., Nakata, K., an Parsons, T.R., eds. Proceedings of Advanced Marine Technology Conferenc (AMTEC), Vol. 5: 57-83. +Ohashi, R., Yamaguchi, A., Matsuno, K., Saito, R., Yamada, N., lijima, A., Shiga, N., an Imai, I. (2013). Interannual changes in the zooplankton community structure o the southeastern Bering Sea shelf during summers of 1994-2009. Deep Se Research Part II: Topical Studies in Oceanography, 94: 44-56. +Okin, G.S., Baker, A.R., Tegen, I., Mahowald, N.M., Dentener, F.J., Duce, R.A., et al (2011). Impacts of atmospheric nutrient deposition on marine productivity: Role of nitrogen, phosphorus, and iron. Global Biogeochemical Cycles, 25, GB2022 doi:10.1029/2010GB003858. +Onabid, M.A. (2011). Improved ocean chlorophyll estimate from remote sensed data The modified blending technique. African Journal of Environmental Science an Technology, 5(9): 732-747. +O'Reilly, J.E., Busch, D.A. (1983). Phytoplankton primary production for th Northwestern Atlantic shelf. Rapports et procés-verbaux des reunions, Consei permanent international pour I'exploration de la mer, 183: 255-268. +O'Reilly, J.E., Maritorena, S., Mitchell, B.G., Siegel, D.A., Carder, K.L., Garver, S.A. Kahru, M., and McClain, C. (1998). Ocean color chlorophyll algorithms fo SeaWiFS. Journal of Geophysical Research, 103 (C11): 24937-24953. +Orhan, I., Sener, B., Atici, T., Brun, R., Perozzo, R., and Tasdemir, D. (2006). Turkis freshwater and marine macrophyte extracts show in vitro antiprotozoal activit and inhibit Fabl, a key enzyme of Plasmodium falciparum fatty acid biosynthesis Phytomedicine, 13(6): 388-393. +Orr, J.C., V. J. Fabry, O. Aumont, L. Bopp, S. C. Doney, R. A. Feely, A. Gnanadesikan, N Gruber, A. Ishida, F. Joos, R. M. Key, K. Lindsay, E. Maier-Reimer, and others (2005). Anthropogenic ocean acidification over the twenty-first century and it impact on calcifying organisms. Nature, 437: 681-686. +Orth, R.J., Carruthers, T.J.B., Dennison, W.C., Duarte, C.M., Fourqurean, J.W. Heck, Jr., K.L., Hughes, A.R., Kendrick, G.A., Kenworthy, W.J., Olyarnik, S. Short, F.T., Waycott, M., and Williams, S.L. (2006). A global crisis for seagras ecosystems. BioScience 56:987-996. +Oschlies, A., Schulz, K.G., Riebesell, U., and Schmittner, A. (2008). Simulated 21 century’s increase in oceanic suboxia by CO2-enhanced biotic carbon export Global Biogeochemical Cycles, +22, GB4008, doi:10.1029/2007GB003147. +Passow, U., and Carlson, C.A. (2012). The biological pump in a high CO2 world. Marin Ecology Progress Series, 470: 249-271. +© 2016 United Nations 5 + +Paulmier, A., and Ruiz-Pino, D. (2009). Oxygen minimum zones (OMZs) in the moder ocean. Progress in Oceanography, 80: 113-28. +Pauly, D., and Christensen, V. (1995). Primary production required to sustain globa fisheries. Nature, (374): 255-257. +Paytan, A., and McLaughlin, K. (2007). The oceanic phosphorus cycle. Chemical Reviews 107, 563-576. +Peierls, B.L., Caraco, N.F., Pace, M.L., and Cole, J.J. (1991). Human influence on rive Nitrogen. Nature, 350: 386-87. +Pennington, J.T., Mahoney, K.L., Kuwahara, V.S., Kolber, D.D., Calienes, R., an Chavez, F.P. (2006). Primary production in the eastern tropical Pacific: A review Progress in Oceanography, 69: 285-317. +Pennock, J.R., and Sharp, J.H. (1986). Phytoplankton production in the Delaware Estuary temporal and spatial variability. Marine Ecology Progress Series, 34: 143-155. +Pereira H., Barreira, L., Figueiredo, F., Custodio, L., Vizetto-Duarte, C., Polo, C., ReSek, E. Engelen, A., and Varela, J. (2012). Polyunsaturated Fatty Acids of Marin Macroalgae: Potential for Nutritional and Pharmaceutical Applications. Marin Drugs, 10(9): 1920-1935. +Pernetta, J.C. (1993). Mangrove forests, climate change and sea level rise: hydrologica influences on community structure and survival, with examples from the Indo- +West Pacific. Marine Conservation Development Report, International Union fo Conservation of Nature, Gland, Switzerland. +Perrings, C., Naeem, S., Ahrestani, F., Bunker, D.E., Burkill, P., et al. (2010). Ecosyste services for 2020. Science, 330: 323-324. +Perry, M.J. (1986). Assessing marine primary production from space. BioScience, 36 (7) 461-467. +PICES. 2004. Marine Ecosystems of the North Pacific. PICES Special Publication 1, 280 p. +Pikitch, E.K., Santora, C., Babcock, E.A., Bakun, A., Bonfil, R., et al. (2004). Ecosystem based fishery management. Science, 305: 346-347. +Pilskaln, C.H., Paduan, J.B., Chavez, F.P., Anderson, R.Y., and Berelson, W.M. (1996) Carbon export and regeneration in the coastal upwelling system of Montere Bay, central California. Journal of Marine Research, 54: 1149-1178. +Pingree, R.D., Pugh, P.R., Holligan, P.M., and Forster, G.R. (1975). Summe phytoplankton blooms and red tides along tidal fronts in the approaches to th English Channel. Nature, 258: 672-677. +Pitt, K.A., Welsh, D.T., Condon, R.H. (2009). Influence of jellyfish blooms on carbon nitrogen and phosphorus cycling and plankton production. Hydrobiologia, 61 (1): 133-149. +© 2016 United Nations 5 + +Platt, T., and Sathyendranath, S. (1993). Estimators of primary production for th interpretation of remotely sensed data on ocean color. Journal of Geophysica Research, 98: 14561-14576. +Platt, T., Fuentes-Yaco, C., and Frank, K. (2003). Marine ecology: Spring algal bloom an larval fish survival. Nature, 423: 398-399. +Platt, T., Hoepffner, N., Stuart, V., and Brown, C., eds. (2008). Why ocean colour? Th Societal benefits of ocean-colour technology. IOCCG Report, No. 7, 141 pp. +Platt, T., White, III, G.N., Zhai, L., Sathyendranath, S., and Roy, S. (2009). The phenolog of phytoplankton blooms: Ecosystem indicators from remote sensing. Ecologica Modelling, 220 (21): 3057-3069. +Polovina, J.J., Howell, E.A., and Abecassis, M. (2008). The ocean’s least productiv waters are expanding. Geophysical Research Letters, 35, L03618 http://dx.doi.org/10.1029/2007GL031745 +Polovina, J.J., Dunne, J.P., Woodworth, P.A., Howell, E.A. (2011). Projected expansion o the subtropical biome and contraction of the temperate and equatoria upwelling biomes in the North Pacific under global warming. /CES Journal o Marine Science, 68 (6), 986-995. +Polovina, J.J., and Woodworth, P.A. (2012). Declines in phytoplankton cell size in th subtropical oceans estimated from satellite remotely-sensed temperature an chlorophyll, 1998-2007. Deep Sea Research Part II: Topical Studies i Oceanography, 77-80, 82-88. +Pomeroy, L.R., Williams, P.J.leB., Azam, F., and Hobbie, J.E. (2007). The microbial loop Oceanography, 20 (2): 28-33. +Powell, T.M., (1989). Physical and biological scales of variability in lakes, estuaries, an the coastal ocean, p. 157-180. In Roughgarden, J., May, R.M., and Levin, S.A. eds. Perspectives in Ecological Theory, Princeton University Press, Princeton, N.J. +Plus, M., Deslous-Paoli, J-M., Auby, I., and Dagault, F. (2001). Factors influencing primar production of seagrass beds (Zostera noltii Hornem.) in the Thau Lagoon (Frenc Mediterranean coast). Journal of Experimental Marine Biology and Ecology 259:63-84. +Primeau, F.W., and Holzer, M. (2006). The Ocean’s Memory of the Atmosphere Residence-Time and Ventilation-Rate Distributions of Water Masses. Journal o Physical Oceanography, 36: 1439-1456. +Purcell, J.E., (2012). Jellyfish and ctenophore blooms coincide with human proliferation and environmental perturbations. Annual Reviews, Marine Science, 4: 209-235. +Rabalais, N.N., Turner, R.E., Diaz, R.J., and Justic, D. (2009). Global change an eutrophication of coastal waters. /CES Journal of Marine Science, 66: 1528-1537. +© 2016 United Nations 5 + +Rahmstorf, S., Foster, G., and Cazenave, A. (2012). Comparing climate projections t observations up to 2011. Environmental Research Letters, 7: 1-5. +Raven, J.A. (1998). Small is beautiful: The picophytoplankton. Functional Ecology, 12 503-513. +Redfield, A.C., Ketchum, B.H., and Richards, F.A. (1963). The influence of organisms o the composition of sea-water. In The Sea, vol. 2, M.N. Hill (ed.), pp. 26-77 Interscience, New York. +Richardson, A.J. (2008). In hot water: zooplankton and climate change. /CES Journal o Marine Science, 65, 279-295. +Richardson, A.J., Bakun, A., Hays, G.C., Gibbons, M.J. (2009). The jellyfish joyride: causes consequences and management responses to a more gelatinous future. Trend in Ecology and Evolution 24(6), 312-322. +Richardson, A.J., and Shoeman, D.S. (2004). Climate impact on plankton ecosystems i the Northeast Atlantic. Science, 305: 1609-1612. +Rockstram, J., Steffen, W., Noone, K., Persson, A., Chapin, F.S. IIl., Lambin, E. Lenton, T.M., Scheffer, M., Folke, C., Schellnhuber, H.J., Nykvist, B., de Wit, B.A. Hughes, T., van der Leeuw, S., Rodhe, H., Sorlin, S., Snyder, P.K., Coastanza, R. Svedin, U., Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J. Walker, B., Liverman, D., Richardson, K., Crutzen, P., Foley, J. (2009). Planetar Boundaries: Exploring the Safe Operating Space for Humanity. Ecology an Society, 14: 32. [online] URL http://www.ecologyandsociety.org/vol14/iss2/art32/. +Rockstrom, J., Steffen, W. Noone, K., Persson, A., Chapin, F.S. Ill, Lambin, E.F. Lenton, T.M., Scheffer, M., Folke, C., Schellnhuber, H.J., Nykvist, B., de Wit, C.A. Hughes, T., van der Leeuw, S., Rodhe, H., Sdrlin, S., Snyder, P.K., Costanza, R. Svedin, U., Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J. Walker, B.H., Liverman, D., Richardson, K., Crutzen, P., and Foley, J.A. (2009). safe operating space for humanity. Nature 461:472-475. +Roelfsema, C.M., Phinn, S.R., Udy, N., and Maxwell, P. (2009). An Integrated Field an Remote Sensing Approach for Mapping Seagrass Cover, Moreton Bay, Australia Spatial Science, 54 (1): 45-62. +Ross, M.S., Ruiz, P.L., Telesnicki, G.J., and Meeder, J.F. (2001). Estimating above-groun biomass and production in mangrove communities of Biscayne National Park Florida (U.S.A.). Wetlands Ecology and Management, 9: 27-37. +Rost, B., Zondervan, I., and Wolf-Gladrow, D. (2008). Sensitivity of phytoplankton t future changes in ocean carbonate chemistry: current knowledge, contradiction and research directions. Marine Ecology Progress Series, 373: 227-237. +Rousseaux, C.S., and Gregg, W.W. (2014). Interannual variations in phytoplankto primary production at a global scale. Remote Sensing, 6: 1-19. +© 2016 United Nations 5 + +Rudnick, D.L., Davis, R.E., Eriksen, C.C., Fratantoni, D.M., and Perry, M.J. (2004) Underwater gliders for ocean research. Marine Technology Society Journal 38(1): 48-59. +Ruttenberg, K.C. (2004). The global phosphorus cycle. In: Holland, H.D. an Turekian, K.K., eds. Treatise on Geochemistry. Elsevier, New York, NY, pp. 585 643. +Rykaczewski, R.R., and Dunne, J.P. (2011). A measured look at ocean chlorophyll trends Nature, 472: E5-E6. +Saba, V.S., Friedrichs, M.A.M., Carr, M-E., Antoine, D., Armstrong, R.A., et al. (2010) Challenges of modeling depth-integrated marine primary productivity ove multiple decades: A case study at BATS and HOT. Global Biogeochemical Cycles 24, GB3020, doi:10.1029/2009GB003655. +Saba, V.S., Friedrichs, M.A.M., Antoine, D., Armstrong, R.A., Asanuma, I. Behrenfeld, M.J., Ciotti, A.M., Dowell, M., Hoepffner, N., Hyde, K.J.W. Ishizaka, J., Kameda, T., Marra, J., Mélin, F., Morel, A., O’Reilly, J., Scardi, M. Smith W.O., Jr., Smyth, T.J., Tang, S., Uitz, J., Waters, K., and Westberry, T.K (2011). An evaluation of ocean color model estimates of marine primar productivity in coastal and pelagic regions across the globe. Biogeosciences, 8 489-503. +Sadeghi, A., Dinter, T., Vountas, M., Taylor, B., Altenburg-Soppa, M., and Bracher, A (2012). Remote sensing of coccolithophore blooms in selected oceanic region using the PhytoDOAS method applied to hyper-spectral satellite data Biogeosciences, 9: 2127-2143. +Sakamoto, C.M., et al. (2004). Influence of Rossby waves on nutrient dynamics and th plankton community structure in the North Pacific subtropical gyre. Journal o Geophysical Research, 109, doi:10.1029/2003JC001976. +Sakshaug, E. (2004). Primary and secondary production in the Arctic Seas. In The Organi Carbon Cycle in the Arctic Ocean, Stein, R., and MacDonald, R.W., eds. pp. 57-81 Springer, Berlin. +Santos, A.L., Oliveira, V., Baptista, |., Henriques, I., Gomes, N.C.M., Almeida, A. Correia, A., and Cunha, A. (2012a). Effects of UV-B radiation on the structurea and physiological diversity of bacterioneuston and bacterioplankton. Applied an Environmental Microbiology, 78: 2066-2069. +Sarmento, H., Montoya, J.M., Vazquez-Dominguez, E., Vaqué, D., and Gasol, J.M. (2010) Warming effects on marine microbial food web processes: how far can we g when it comes to predictions? Philosophical Transactions of the Royal Society B 365: 2137-2149. +© 2016 United Nations 5 + +Sathyendranath, S., ed. (2000). Remote sensing of ocean colour in coastal and othe optically complex waters. Reports of the International Ocean-Colou Coordinating Group, No. 3, 140 p., Dartmouth, Canada. +Sathyendranath, S., Ahanhanzo, J., Bernard, S., Byfield, V., Delaney, L., and others (2010). ChloroGIN: Use of satellite and in situ data in support of ecosystem based management of marine resources in Proceedings of OceanObs’09 Sustained Ocean Observations and Information for Society (Vol. 2), Venice, Italy 21-25 September 2009, Hall, J., Harrison, D.E., & Stammer, D., eds. ES Publication WPP-306. +Scavia, D., and Bricker, S.B. (2006). Coastal eutrophication assessment in the Unite States. Biogeochemistry, 79: 187-208. +Schlitzer, R., Usbeck, R., and Fischer, G. (2003). Inverse modeling of particulate organi carbon fluxes in the South Atlantic. In: The South Atlantic in the Late Quaternary Reconstruction of material budget and current systems. Wefer, G., Mulitza, S. and Ratmeyer, V., eds. Springer, Berlin, Heidelberg, New York, pp. 1-19. +Schlitzer, R., Monfray, P., Hoepffner, N., Fischer, G., Gruber, N., Lampitt, R., Lévy, M. Schmidt, A.L., Wysmyk, J.K.C., Craig, S.E., and Lotze, H.K. (2012). Regional-scal effects of eutrophication on ecosystem structure and services of seagrass beds Limnology and Oceanography. 57(5): 1389-1402. +Schmidt, A.L., Wysmyk, J.K.C., Craig, S.E., and Lotze, H.K. (2012). Regional-scale effects o eutrophication on ecosystem structure and services of seagrass beds. Limnolog and Oceanography. 57(5): 1389-1402. +Schneider, B., Bopp, L., Gehlen, M., Segschneider, J., Frolicher, T.L., Cadule, P. Friedlingstein, P., Doney, S.C., Behrenfeld, M.J., and Joos, F. (2008). Climate induced interannual variability of marine primary and export production in thre global coupled climate carbon cycle models. Biogeosciences, 5: 597-614. +Schulz, M., Prospero, J.M., Baker, A.R., Dentener, F., et al. (2012). Atmospheric transpor and deposition of mineral dust to the ocean: implications for research needs Environmental Science & Technology, 46: 10,390-10,404. +Seitzinger, S.P., Harrison, J.A., Dumont, E., Beusen, A.H.W., Bouwman A.F. (2005) Sources and delivery of carbon, nitrogen, and phosphorus to the coastal zone An overview of the Global Nutrient Export from Watersheds (NEWS) models an their application, Global Biogeochemical Cycles 19, GB4S01, doi:10.1029/2005GB002606. +Seitzinger, S.P., Mayorga, E., Bouwman, A.F., et al. (2010). Global nutrient export: scenario analysis of past and future trends. Global Biogeochemical Cycles, 23 GBOAO8, doi: 10.1029/2009GB003587. +Sharma, K.K., Schuhmann, H., and Schenk, P.M. (2012). High lipid induction i microalgae for biodiesel production. Energies, 5: 1532-1553. +© 2016 United Nations 6 + +Sharpies, J., Middelburg, J.J., Fennel, K., Jickells, T.D. (2013), Riverine delivery o nutrients and carbon to the oceans, Nature, 504:61-70 doi:10.1038/nature12857. +Sherman, K., and Hempel, G. (2009). The UNEP Large Marine Ecosystem Report: perspective on changing conditions in LMEs of the world's Regional Seas. UNE Regional Seas Reports and Studies, 182. United Nations Environmenta Programme, Nairobi, 872 pp. +Sherman, K., Alexander, L.M., and Gold, B.D., eds. (1993). Large Marine Ecosystems Stress, Mitigation, and Sustainability, AAAS Press, Washington, D.C., 376 pp. +Sherman, K., Smith, W., Morse, W., Berman, M., Green, J., and Ejsymont, L. (1984) Spawning strategies of fishes in relation to circulation, phytoplankto production, and pulses in zooplankton off the northeastern United States Marine Ecology Progress Series, 18: 1-19. +Shi, D., Xu, Y., and Morel, F.M.M. (2009). Effects of the pH/pCO2 control method o medium chemistry and phytoplankton growth. Biogeosciences, 6: 1199-1207. +Shi, D., Xu, Y., Hopkinson, B.M., Morel, F.M.M. (2010). Effect of ocean acidification o iron availability to marine phytoplankton. Science, 327(5966): 676-679. +Shi, D., Kranz, S.A., Kim, J-M., and Morel, F.M.M. (2012). Ocean acidification slow nitrogen fixation and growth in the dominant diazotroph Trichodesmium unde low-iron conditions. Proceedings of the National Academy of Sciences, 109 (45) E3094-100. +Short, F.T., and Neckles, H.H. (1999). The effects of global climate change on seagrasses Aquatic Botany, 63: 169-196. +Shurin, J.G., Gruner, D.S., and Hillebrand, H. (2006). All wet or dried up? Real difference between aquatic and terrestrial food webs. Proceedings of the Royal Society B 273: 1-9. +Siegel, D.A., Behrenfeld, M.J., Maritorena, S., McClain, C.R., Antoine, D., and others. +(2013). Regional to global assessments of phytoplankton dynamics from th SeaWiFS mission. Remote Sensing of Environment 135: 77-91. +Siegel, D.A., Doney, S.C., and Yoder, J.A. (2002). The North Atlantic spring phytoplankto bloom and Sverdrup’s critical depth hypothesis. Science, 296: 730- 733. +Silliman, B.R., Grosholz, T., and Bertness, M.D. (2009). Salt marshes under global siege Pages 103-114 in Silliman, B.R., Grosholz, T., and Bertness, M.D., eds. Huma impacts on salt marshes: a global perspective. University of California Press Berkeley, California, USA. +Silva, J., Sharon, Y., Santos, R., and Beer, S. (2009). Measuring seagrass photosynthesis methods and applications. Aquatic Biology, 7: 127-141. +© 2016 United Nations 6 + +Slomp, C.P. (2011). Phosphorus cycling in the estuarine and coastal zones: sources sinks, and transformations. In: Wolanski, E., and McLusky, D.S., eds. Treatise o Estuarine and Coastal Science, Vol 5, pp. 201-229. Waltham: Academic Press. +Smetacek, V. (1999). Diatoms and the ocean carbon cycle. Protist, 150 (1): 25-32. +Smith, R.C., Baker, K.S., Dierssen, H.M., Stammerjohn, S.E., and Vernet, M. (2001) Variability of primary production in an Antarctic marine ecosystem as estimate using a multi-scale sampling strategy. American Zoology, 41: 40-56. +Smith, S.V. (1981). Marine macrophytes as a global carbon sink. Science, 211: 838-840. +Smith, Jr., W.O., and Demaster, D.J. (1996). Phytoplankton biomass and productivity i the Amazon River plume: correlation with seasonal river discharge. Continenta Shelf Research, 16 (3): 291-319. +Sohm, J.A., Webb, E.A., and Capone, D.G. (2011). Emerging patterns of marine nitroge fixation. Nature Reviews: Microbiology, 9: 499-508. +Son, N-T., and Chen, C-F. (2013). Remote sensing of mangrove forests in Centra America. SPIE Newsroom, doi: 10.1117/2.1201304.00477 (http://spie.org/x93599.xml) +Stal, L.J. (2009). Is the distribution of nitrogen-fixing cyanobacteria in the oceans relate to temperature? Environmental Microbiology, 11 (7): 1632-45. +Steele, J.H. (1985). A comparison of terrestrial and marine ecological systems. Nature 313, 355-358. +Steeman-Nielsen, E. (1963). Productivity, definition and measurement. In Hill, M.W. ed The Sea, vol. 1, pp 129-164, New York, John Wiley. +Steinacher, M., Joos, F., Frélicher, T.L., Bopp, L., Cadule, P., Cocco, V., Doney, S.C. Gehlen, M., Lindsay, K., Moore, J.K., Schneider, B., and Segschneider, J. (2010) Projected 21st century decrease in marine productivity: a multi-model analysis Biogeosciences, 7: 979-1005. +Steinberg, D.K., Carlson, C.A., Bates, N.R., Johnson, R.J., Michaels, A.F., and Knap, A.H (2001). Overview of the US JGOFS Bermuda Atlantic Time-series Study (BATS): decade-scale look at ocean biology and biogeochemistry. Deep-Sea Research II 48: 1405-1447. +Steinberg, D.K., Lomas, M.W., Cope, J.S., (2012). Long-term increase i mesozooplankton biomass in the Sargasso Sea: Linkage to climate an implications for food web dynamics and biogeochemical cycling. Globa Biogeochemical Cycles 26 (1), GB1004. +Stier, P., Feichter, J., Kinne, S., Kloster, S., Vignati, E., Wilson, J., Ganzeveld, L., Tegen, I. Werner, M., Balkanski, Y., Schulz, M., Boucher, O., Minikin, A., and Petzold, A (2005). The aerosol-climate model ECHAM5-HAM. Atmospheric Chemistry an Physics, 5: 1125-1156. +© 2016 United Nations 6 + +Strous, M., Pelletier, E., Mangenot, S., et al. (2006). Deciphering the evolution an metabolism of an anammox bacterium from a community genome. Nature, 440 790-794. +Subramaniam, A., Yager, P.L., Carpenter, E.J., Mahaffey, C., Bjorkman, K., Cooley, S. Kustka, A.B., Montoya, J.P., Safiudo-Wilhelmy, S.A., Shipe, R., and Capone, D.G (2008). Amazon River enhances diazotrophy and carbon sequestration in th tropical North Atlantic Ocean. PNAS, +105 (3): 10460-10465. +Suntharalingam, P., Buitenhuis, E., Le Quéré, C., Dentener, F., Nevison, C., Butler, J.H. Bange, H.W., and Forster, G. (2012). Quantifying the impact of anthropogeni nitrogen deposition on oceanic nitrous oxide, Geophysical Research Letters, 39 LO7605, doi:10.1029/2011GL050778. +Suikkanen, S., Pulina, S., Engstrém-Ost, J., Lehtiniemi, M., Lehtinen, S., et al. (2013 Climate Change and Eutrophication Induced Shifts in Northern Summer Plankto Communities. PLoS ON 8 (6): 66475. doi:10.1371/journal.pone.0066475. +Taalas, P., Kaurola, J., Kylling, A., Shindell, D., Sausen, R., Dameris, M., Grewe, V. Herman, J., Damski, J. and Steil, B. (2000). The impact of greenhouse gases an halogenated species on future solar UV radiation doses. Geophysical Researc Letters, 27: 1127-1130. +Taucher, J. and Oschlies, A. (2011). Can we predict the direction of marine primar production change under global warming? Geophysical Research Letters, 38 LO2603. +Taucher, J., Schulz, K.G., Dittmar, T., Sommer, U., Oschlies, A., and Riebesell, U. (2012 Enhanced carbon overconsumption in response to increasing temperature during a mesocosm experiment. Biogeosciences, 9: 3479-3514. +Teira, E., Mourifio, B., Marafion, E., Pérez, V., Pazo, M.J., Serret, P., de Armas, D. Escdnez, J., Woodward, E.M.S., and Fernandez, E. (2005). Variability o chlorophyll and primary production in the eastern north Atlantic subtropica gyre: potential factors affecting phytoplankton activity. Deep-Sea Research, 52 569-588. +Testor, P., Meyers, G., Pattiaratchi, C., Bachmayer, R., Hayes, D., Pouliquen, S., et al (2010). Gliders as a component of future observing systems in Proceedings o OceanObs’09: Sustained Ocean Observations and Information for Society (Vol 2), Hall, J., Harrison, D.E., and Stammer, D., eds. ESA Publication WPP-306 doi:10.5270/OceanObs09. +Teuten, E.L., Saquing, J.M., Knappe, D.R.U., Barlaz, M.A., Jonsson, S., Bjorn, A. Rowland, S.J., Thompson, R.C., Galloway, T.S., Yamashita, R., Ochi, D., Watanuki Y., Moore, C., Viet, P.H., Tana, T.S., Prudente, M., Boonyatumanond, R., Zakaria M.P., Akkhavong, K., Ogata, Y., Hirai, H., lwasa, S., Mizukawa, K., Hagino, Y., +© 2016 United Nations 6 + +Imamura, A., Saha, M., Takada, H. (2009). Transport and release of chemical from plastics to the environment and to wildlife. Philosophical Transactions o the Royal Society of London B: Biological Science, 364: 2027-2045. +Thamdrup, B., Dalsgaard, T., Jensen, M.M., et al. (2006). Anaerobic ammoniu oxidation in the oxygen-deficient waters off northern Chile. Limnology an Oceanography, 51: 2145-2156. +Tian, T., Wei, H., Su, J., and Chung, C. (2005). Simulations of annual cycle o phytoplankton production and the utilization of nitrogen in the Yellow Sea Journal of Oceanography, 61: 343-357. +Titus, J.G., Anderson, K.E., Cahoon, D.R., Gesch, D.B., Gill, S.K, Gutierrez, B.T. Thieler, E.R., and Williams, S.J. (2009). Coastal sensitivity to sea level rise: A Focu on the Mid-Atlantic Region. U.S. Climate Change Science Program, Synthesis an Assessment Product 4.1. +Tremblay, J.E., and Gagnon, J. (2009). The effect of irradiance and nutrient supply on th productivity of Arctic waters: A perspective on climate change, pp. 73-94, I Influence of Climate Change on the Changing Arctic and Sub-Arctic Conditions Springer, New York. +Trilla, G.G., Pratolongo, P., Beget, M.E., Kandus, P, Marcovecchio, J., and Di Bella, C (2013). Relating biophysical parameters of coastal marshes to hyperspectra reflectance data in the Bahia Blanca Estuary, Argentina. Journal of Coasta Research, 29 (1): 231-238. +Twilley, R.R., Chen, R.H., and Hargis, T. (1992). Carbon sinks in mangroves and thei implications to carbon budget of tropical coastal ecosystems. Water Air and Soi Pollution, 64 (1): 265-265. +Ueyama, R., and Monger, B.C. 2005. Wind-induced modulation of seasona phytoplankton blooms in the North Atlantic derived from satellite observations Limnology and Oceanography, 50: 1820- 1829. +Ulloa, O., Canfieldb, D.E., DeLong, E.F., Letelier, R.M., and Stewart, F.J. (2012). Microbia oceanography of anoxic oxygen minimum zones. Proceedings of the Nationa Academy of Sciences (USA), 109 (40): 15996-16003. +UNEP-WCMC. (2006). In the front line: Shoreline protection and other ecosystem service from mangroves and coral reefs. UNEP-WCMC, Cambridge, UK 33 pp. +UNESCO. (2006). A Coastal Theme for the IGOS Partnership for the Monitoring of ou Environment from Space and from Earth. Paris, 60 pp. +UNESCO. (2008). Filling Gaps in LME Nitrogen Loadings Forecast for 64 LMEs. GEF/LME +UNESCO-IOC. (2012). Requirements for Global Implementation of the Strategic Plan fo Coastal GOOS. GOOS Report 193. Paris: Intergovernmental Oceanographi Commission. (www.ioc- +© 2016 United Nations 6 + +goos.org/index.php?option=com_oe&task=viewDocumentRecord&doclD=7702 lang=en) +Uye, S. (2008). Blooms of the giant jellyfish Nemopilema nomurai: A threat to th fisheries sustainability of the East Asian Marginal Seas. Plankton & Bentho Research, 3: 125-131. +Valiela, |., Bowen, J.L., and York, J.K. (2001). Mangrove forests: One of the world’ threatened major tropical environments. BioScience, 51 (10): 807-815. +Valiela, |., Kinney, E., Culbertson, J., Peacock, E., and Smith, S. (2009). Global losses o mangroves and salt marshes: Magnitudes, causes and consequences. Pp. 107 138 in Global Loss of Coastal Habitats: Rates, Causes, and Consequences, Duarte C. ed. Fundacion BBVA. Bilbao. +Vancoppenolle, M., Bopp, L., Madec, G., Dunne, J., Ilyina, T., Halloran, P.R., an Steiner, N. (2013), Future Arctic Ocean primary productivity from CMIP simulations: Uncertain outcome, but consistent mechanisms, Globa Biogeochemical Cycles, 27, doi:10.1002/gbc.20055. +Van Dolah, F.M. (2000). Marine algal toxins: origins, health effects, and their increase occurrence. Environmental Health Perspectives, 108 (suppl 1): 133-141. +Vantrepotte, V., and Mélin, F. (2009). Temporal variability of 10-year global SeaWiF time-series of phytoplankton chlorophyll a concentration. ICES Journal of Marin Science, 66: 1547-1556. +Vargas, M., Brown, C.W., and Sapiano, M.R.P. (2009). Phenology of marin phytoplankton from satellite ocean color measurements. Geophysical Researc Letters, 36: LO1608, doi: 10.1029/2008GL036006. +Vernet, M., and Smith, R.C. (2007) Measuring and Modeling Primary Production i Marine Pelagic Ecosystems, in Principles and Standards for measuring Primar Production, Fahey, J.T., and Knapp, A.K., eds. Oxford University Press, Ch.9 pp.142-174. +Vernet, M., Martinson, D., lannuzzi, R., Stammerjohn, S., Kozlowski, W., Sines, K. Smith, R.C., and Garibotti, |. (2008). Primary production within the sea-ice zon west of the Antarctic Peninsula: | — Sea ice, summer mixed layer, and irradiance Deep Sea Research Part II: Topical Studies in Oceanography, 55: 2068-2085. +Villar-Argaiz, M., Median-Sanchez, J.M., Bullejos, F.J., Delgado-Molina, J.A., Rui Pérez, O., Navarro, J.C., and Carrillo, P. (2009). UV radiation and phosphoru interact to influence the biochemical composition of phytoplankton. Freshwate Biology, 54, 1233-1245. +Vonthron-Sénécheau, C., Kaiser, M., Devambez, L., Vastel, A., Mussio, I., Rusig, A.M (2011). Antiprotozoal activities of organic extracts from French marin seaweeds. Marine Drugs, +9 (6):922-33. +© 2016 United Nations 6 + +Voss, M., Bange, H.W., Dippner, J.W., Middelburg, J.J., Montoya, J.P., and Ward, B.B (2013). The marine nitrogen cycle: recent discoveries, uncertainties and th potential relevance of climate change. Philosophical Transactions of the Roya Society B, Biological Sciences, 368 (1621): 20130121 doi: 10.1098/rstb.2013.0121. +Walker, B., Holling, C.S., Carpenter, S.R., Kinzig, A. (2004). "Resilience, adaptability an transformability in social-ecological systems". Ecology and Society, 9 (2): 5. +Walsh, P.J., Smith, S.L., Fleming, L.E., Solo-Gabriele, H.M., and Gerwick, W.H., eds (2008). Oceans and Human Health: Risk and Remedies from the Sea. Elsevier, 64 p. +Wang, M., and Overland, J. (2009). An sea ice free summer Arctic within 30 years Geophysical Research Letters, 36, LO7502 http://dx.doi.org/10.1029/2009GL037820. +Ward, B.A., Dutkiewicz, S., Jahn, O., and Follows, M.J. (2012). A size-structured food web model for the global ocean. Limnology and Oceanography, 57(6): 1877 1891. +Ward, B.B. (2012). The Global Nitrogen Cycle. In: Knoll, A.H., Canfield, D.E., an Konhauser, K.O., eds, Fundamentals of Geomicrobiology, Wiley-Blackwell Chichester, UK, pp. 36-48. +Ward, B.B. (2013). How is nitrogen lost? Science, 341: 352 — 352. +Ward, B.B., Capone, D.G., and Zehr, J.P. (2007). What’s New in the Nitrogen Cycle Oceanography, 20 (2): 101-109. +Ward, B.B., Devol, A.H., Rich, J.J., Chang, B.X., Bulow, S.E., Naik, H., Pratihary, A., an Jayakumar, A. (2009). Denitrification as the dominant nitrogen loss process i the Arabian Sea, Nature, 461:78-81. +Ware, D.M. (2000). Aquatic ecosystems: Properties and models. In: Harrison P.J. Parsons T.R., eds. Fisheries oceanography: An integrative approach to fisherie ecology and management. Oxford: Balckwell Science. pp. 161-200. +Ware, D.M. and Thomson, R.E. (2005). Bottom-up ecosystem trophic dynamic determine fish production in the northeast Pacific. Science, 308: 1280-1284. +Waycott, M., Duarte, C.M., Carruthers, T.J.B., Orth, R.J., Dennison, W.C., Olyarnik, S. Calladine, A., Fourqurean, J.W., Heck, K.L., Hughes, A.R., Kendrick, G.A. Kenworthy, W.J., Short, F.T., and Williams, S.L. (2009). Accelerating loss o seagrasses across the globe threatens coastal ecosystems. Proceedings of th National Academy of Sciences USA 106:12377-12381. +Wegner, I.A., Besseling, E., Foekema, E.M., Kamermans, P., and Koelmans, A.A. (2012) Effects of nanopolystyrene on the feeding behavior of the blue mussel (Mytilu edulis L.). Environmental Toxicology and Chemistry, 31 (11): 2490-2497. +© 2016 United Nations 6 + +Wernand, M.R., van der Woerd, H.J., Gieskes, W.W.C. (2013). Trends in ocean colou and chlorophyll concentration from 1889 to 2000, worldwide. PLoS One, 8(6). +Wernberg, T., Russell, B.D., Thomsen, M.S., Gurgel, C.F.D., Bradshaw, C.J.A. Poloczanska, E.S., Connell, $.D. (2011). Seaweed communities in retreat fro ocean warming. Current Biology, +21: 1828-1832. +Westberry, T., Behrenfeld, M.J., Siegel, D.A., and Boss, E. (2008). Carbon based primar productivity modeling with vertically resolved photoacclimation. Globa Biogeochemical Cycles, 22, GB2024, doi:10.1029/2007GB003078. +Wollast, R. (1998). Evaluation and comparison of the global carbon cycle in the coasta zone and the open ocean. In The Sea, Brink, K.H. and A.R. Robinson, eds. 10: 213 253. +Worm, B., Barbier, E.B., Beaumont, N., Duffy, J.E., Folke, C., Halpern, B.S., Jackson, J.B.C. Lotze, H.K., Micheli, F., and Palumbi, S.R. (2006). Impacts of biodiversity loss o ocean ecosystem services. Science, 314: 787-790. +Wright, S.L., Thompson, R.C., and Galloway, T.S. (2013). The physical impacts o microplastics on marine organisms: A review. Environmental Pollution, 178 483e492. +Yusuf, C. (2007). Biodiesel from microalgae. Biotechnology Advances, 25:294-306. +Zehr, J.P., Kudela, R.M. (2011), Nitrogen Cycle of the Open Ocean: From Genes t Ecosystems. Annual Reviews Marine Science, 3: 197-225. +Zhang, M., Ustin, S.L., Rejmankova, E., and Sanderson, E.W. (1997). Monitoring Pacifi coast salt marshes using remote sensing. Ecological Applications, 7 (3): 1039 1053. +Zhai, L., Platt, T., Tang, C., Sathyendranath, S., and Walne, A. (2013). The response o phytoplankton to climate variability associated with the North Atlanti Oscillation. Oceanography, 93: 159-168. +© 2016 United Nations 6 +