13
MARINE ECOLOGY PROGRESS SERIES Mar Ecol Prog Ser Vol. 494: 191–203, 2013 doi: 10.3354/meps10572 Published December 4 INTRODUCTION A central goal of ecology is to understand patterns of abundance and distribution of organisms. Deter- mining the key factors that drive population dynam- ics requires knowledge of demographic inputs (birth and immigration) and losses (death and emigration). In marine systems, the process of larval dispersal adds additional complexity to quantifying population inputs as many marine organisms release dispersive larvae that act as the primary agents coupling birth at one site to immigration at another. Therefore, larval recruitment to coastal populations relies not only on reproductive output of parent populations and post- settlement processes, but also on oceanographic fac- tors that affect the dispersal and delivery of settlers © Inter-Research 2013 · www.int-res.com *Email: [email protected] Spatial differences in larval abundance within the coastal boundary layer impact supply to shoreline habitats Kerry J. Nickols 1,4, *, Seth H. Miller 1 , Brian Gaylord 1,2 , Steven G. Morgan 1,3 , John L. Largier 1,3 1 Bodega Marine Laboratory, University of California at Davis, Bodega Bay, California 94923, USA 2 Department of Evolution and Ecology, University of California at Davis, Davis, California 95616, USA 3 Department of Environmental Science and Policy, University of California at Davis, Davis, California 95616, USA 4 Present address: Hopkins Marine Station, Stanford University, Pacific Grove, California 93950, USA ABSTRACT: Explorations of the dynamics of nearshore regions of the coastal zone are missing from many efforts to understand larval transport and delivery to suitable habitats. Larval distribu- tions in the coastal ocean are variable and depend on physical processes and larval behaviors, leading to biophysical interactions that may increase larval retention nearshore and bolster their return to natal sites. While recent evidence suggests that many larvae are retained within a few kilometers from shore, few studies incorporate measurements sufficiently close to shore to plausi- bly assess supply to the shoreline benthos. We measured cross-shore distributions of larvae of benthic crustaceans between 250 and 1100 m from shore (i.e. just beyond the surf zone) within the coastal boundary layer (CBL) — a region of reduced alongshore flow — and simultaneously quan- tified a suite of physical factors that may influence larval distributions. We found high larval abun- dance within the CBL, with a peak at 850 m from shore, and a decrease in abundance along the shoreward edge of the sampled transect. We also found distinctly different larval assemblages at outer stations within the CBL, as compared to inner stations that are more influenced by shoreline dynamics. These patterns persisted across sample dates, suggesting that the spatial structure of nearshore larval assemblages is at least somewhat robust to temporal changes in physical condi- tions. Thus, while larval abundance appears to be high within the CBL, larvae appear to be sparse within the narrow band of water adjacent to the surf zone. Low larval supply adjacent to suitable habitats has important implications for the coupling of supply and recruitment, and resulting dynamics of shoreline populations. KEY WORDS: Dispersal · Invertebrate larvae · Retention · Nearshore · Transport Resale or republication not permitted without written consent of the publisher FREE REE ACCESS CCESS

Spatial differences in larval abundance within the coastal boundary

Embed Size (px)

Citation preview

  • MARINE ECOLOGY PROGRESS SERIESMar Ecol Prog Ser

    Vol. 494: 191203, 2013doi: 10.3354/meps10572

    Published December 4

    INTRODUCTION

    A central goal of ecology is to understand patternsof abundance and distribution of organisms. Deter-mining the key factors that drive population dynam-ics requires knowledge of demographic inputs (birthand immigration) and losses (death and emigration).In marine systems, the process of larval dispersal

    adds additional complexity to quantifying populationinputs as many marine organisms release dispersivelarvae that act as the primary agents coupling birth atone site to immigration at another. Therefore, larvalrecruitment to coastal populations relies not only onreproductive output of parent populations and post-settlement processes, but also on oceanographic fac-tors that affect the dispersal and delivery of settlers

    Inter-Research 2013 www.int-res.com*Email: [email protected]

    Spatial differences in larval abundance withinthe coastal boundary layer impact supply to

    shoreline habitats

    Kerry J. Nickols1,4,*, Seth H. Miller1, Brian Gaylord1,2, Steven G. Morgan1,3, John L. Largier1,3

    1Bodega Marine Laboratory, University of California at Davis, Bodega Bay, California 94923, USA2Department of Evolution and Ecology, University of California at Davis, Davis, California 95616, USA

    3Department of Environmental Science and Policy, University of California at Davis, Davis, California 95616, USA

    4Present address: Hopkins Marine Station, Stanford University, Pacific Grove, California 93950, USA

    ABSTRACT: Explorations of the dynamics of nearshore regions of the coastal zone are missingfrom many efforts to understand larval transport and delivery to suitable habitats. Larval distribu-tions in the coastal ocean are variable and depend on physical processes and larval behaviors,leading to biophysical interactions that may increase larval retention nearshore and bolster theirreturn to natal sites. While recent evidence suggests that many larvae are retained within a fewkilometers from shore, few studies incorporate measurements sufficiently close to shore to plausi-bly assess supply to the shoreline benthos. We measured cross-shore distributions of larvae ofbenthic crustaceans between 250 and 1100 m from shore (i.e. just beyond the surf zone) within thecoastal boundary layer (CBL) a region of reduced alongshore flow and simultaneously quan-tified a suite of physical factors that may influence larval distributions. We found high larval abun-dance within the CBL, with a peak at 850 m from shore, and a decrease in abundance along theshoreward edge of the sampled transect. We also found distinctly different larval assemblages atouter stations within the CBL, as compared to inner stations that are more influenced by shorelinedynamics. These patterns persisted across sample dates, suggesting that the spatial structure ofnearshore larval assemblages is at least somewhat robust to temporal changes in physical condi-tions. Thus, while larval abundance appears to be high within the CBL, larvae appear to be sparsewithin the narrow band of water adjacent to the surf zone. Low larval supply adjacent to suitablehabitats has important implications for the coupling of supply and recruitment, and resultingdynamics of shoreline populations.

    KEY WORDS: Dispersal Invertebrate larvae Retention Nearshore Transport

    Resale or republication not permitted without written consent of the publisher

    FREEREE ACCESSCCESS

  • Mar Ecol Prog Ser 494: 191203, 2013

    (Cowen 1985, Gaines et al. 1985, Roughgarden et al.1988, Gaylord & Gaines 2000, Morgan 2001, Under-wood & Keough 2001, Largier 2003, Prairie et al.2012). While larval recruitment is a critical determi-nant of population structure (Gaines & Roughgarden1985, Underwood & Fairweather 1989, Menge et al.2004) and larval delivery has long been recognizedfor its role in driving population dynamics (Thorson1946), larval transport pathways and connectivity aredifficult to quantify due to the small size of larvae andthe difficulty of tracking them (Levin 2006). Thesedifficulties are exacerbated by a lack of understand-ing of how larval supply varies over time and space.

    Relatively few studies have measured larval supplyand settlement concurrently in such a way as to ex -plicitly link them. Those that have done so havereached mixed conclusions: some found supply andsettlement to be coupled (e.g. Gaines et al. 1985,Bertness et al. 1992, Gaines & Bertness 1992, Dudaset al. 2009) while others did not (e.g. Yoshioka 1982,McCulloch & Shanks 2003, Rilov et al. 2008). Suchdifferences may be due to the magnitude of recruit-ment, with larval supply acting as a strong predictorof adult dynamics in regions where recruitment islimited (Connell 1985), but less so in areas where re -cruitment is high. Connections between supply andsettlement or recruitment may be easier to assess inother systems; for instance those involving dispersalof macroalgal spores (see, e.g. Reed et al. 1988, Gay-lord et al. 2002, 2004, 2006, 2012, Reed et al. 2006).

    Much of the west coast of North America is recruit-ment-limited to one degree or another, such that lar-val supply is a critical determinant of populationdynamics in many species. This feature derives fromthe fact that the west coast sits within a major up -welling region. During times of strong equatorwardwinds, the predominant currents flow equatorwardand surface waters move offshore. The potential forupwelling waters to move larvae offshore has beenwidely recognized (Yoshioka 1982, Roughgarden etal. 1988, Botsford et al. 1994), and is consistent withobservations that larval settlement and supply in persistent upwelling regions is higher during relax-ation events when wind speeds decrease or reversedirections (Farrell et al. 1991, Botsford 2001). In -creases in settlement and supply during relaxationevents can also result from poleward advection of lar-vae (Wing et al. 1995a,b). More recent work hasshown the persistence of sequential larval stages innearshore plankton during upwelling conditions,indicating that many taxa are not swept offshore inupwelling regions (Morgan et al. 2009b, Shanks &Shearman 2009). Larvae appear to be able to at least

    partially avoid offshore transport associated withupwelling through physical and behavioral mecha-nisms. Nearshore retention zones arising from topo-graphic effects on coastal circulation have been ob -served (Graham & Largier 1997, Wing et al. 1998,Roughan et al. 2005), and are associated with higherlarval abundances and settlement (Mace & Morgan2006, Morgan et al. 2009a, 2011, 2012). Avoidance ofsurface waters by larvae can favor retention and de -crease offshore transport (Morgan et al. 2009b,c,Shanks & Shearman 2009, Morgan & Fisher 2010,Morgan et al. 2012) and there is mounting evidencethat larval concentrations are high close to shore,even in areas of strong upwelling that are tradition-ally viewed as being recruitment-limited. Here weinvestigate this phenomenon closer to shore to see ifhigh abundances extend over the inner shelf andinward to the shoreline.

    Nearshore processes play an important role in lar-val ecology, and may be relevant for a significantportion of pelagic larval durations. A number of re -cent studies measured larval abundance in cross-shore transects and found increases in abundancecloser to shore in a range of invertebrates and fishes(Borges et al. 2007, Tapia & Pineda 2007, Morgan etal. 2009b,c, Shanks & Shearman 2009). For example,Morgan et al. (2009b) measured larval abundance ofbenthic crustaceans from 1 km from shore out to theshelf break (30 km offshore) along the open coast ofnorthern California and found that the highest larvalabundances were within 3 km from shore. The com-bination of larval behaviors (e.g. swimming and ver-tical migration) and nearshore processes may in -crease retention of larvae close to shore and to theirnatal site. Such retention is consistent with evidencefrom a range of species and systems that shows self-recruitment is higher and dispersal distances smallerthan previously thought (Swearer et al. 2002, Levin2006, Shanks 2009), and further emphasizes theimportance of understanding the role of nearshoreprocesses in larval supply and population dynamics.

    A number of nearshore processes may reducescales of dispersal. Adjacent to the shore within thesurf zone, rip tides can create recirculation zones(MacMahan et al. 2010) and onshore wave transportcan lead to accumulation of water-borne material(Monismith 2004, McPhee-Shaw et al. 2011). There isalso a region termed the coastal boundary layer(CBL) that extends beyond the surf zone and is char-acterized by reduced speeds (Nickols et al. 2012). Inparticular, average alongshore velocities in the CBLare an order of magnitude larger than cross-shorevelocities (Lentz et al. 1999, Gaylord et al. 2007), and

    192

  • Nickols et al.: Larvae in the coastal boundary layer

    alongshore velocities increase strongly with distancefrom shore until reaching a free-stream value off-shore (Nickols et al. 2012). Such decreased velocitiesprovide another potential mechanism for reducingscales of dispersal in coastal populations (Nickols etal. 2012, Nickols et al. unpubl. data), but require thatlarvae spend sufficient time within the CBL forreduced flow to influence net transport. Because pre-vious studies of coastal larval distributions did notextend into the CBL or only just entered the CBL(McQuaid & Phillips 2000, Morgan et al. 2009b,c,Shanks & Shearman 2009), or sampled over an insuf-ficient temporal scale to fully characterize patterns(Tapia & Pineda 2007), the general role of the CBL ininfluencing patterns of larval transport and supplyremains unknown.

    The goals of this study were therefore to addressthe following questions: (1) what is the spatial patternof larval abundance within the CBL, (2) are there differences in larval assemblages close to shore (i.e.are different larvae found inshore versus offshorewithin the CBL), and (3) do time-dependent or space- de pendent physical processes tend to dictate vari-ability in larval abundance? We recognize in target-ing these goals that it is not possible to definitivelyascribe particular mechanisms to observed patterns;rather, our aim is to present the first description of thespatio-temporal distribution of larval assemblageswithin the CBL, which is critical for understandingthe potential for nearshore transport processes toimpact larval supply to shoreline habitats. We addi-tionally hope that this study will also inform themethodological question of where supply should bemeasured to best address relationships between set-tlement and supply when exploring questions aboutmarine population dynamics, particularly for recruit-ment- limited regions.

    MATERIALS AND METHODS

    Study system and species

    This study was conducted along the open coast innorthern California, USA, near Bodega Head, Cali-fornia (Fig. 1), a region characterized by strong sea-sonal upwelling during spring and summer thatdrives prevailing currents equatorward and pushessurface waters offshore (Winant et al. 1987, Largier etal. 1993). When the winds weaken or reverse direc-tion (inducing a relaxation event), currents movepoleward, often responding within a day or less invery nearshore regions (Send et al. 1987). Inner shelf

    currents, observed previously on the scale of kilome-ters, are slower than currents farther offshore, have ahigher tendency to move poleward (Kaplan et al.2005), and are associated with increases in inverte-brate settlement (Wing et al. 1995a,b). Benthic crus -ta cean larvae can be present during both relaxationand upwelling conditions within areas of topographicretention and along the open coast, and many can beretained in areas within 1 to 3 km of the shore via acombination of physical and behavioral mechanisms(Morgan et al. 2009b, Morgan & Fisher 2010, Morganet al. 2012). The present study focused on waterswithin 1 km of the shore.

    Our efforts focused on larvae of benthic crusta -ceans (primarily barnacles and crabs), which are thebest-studied meroplankton in this region. From priorwork, it is known that larval abundance of benthiccrustaceans peaks during the spring and summer,coinciding with the upwelling season. Barnacles moltthrough 6 larval stages (nauplii) and a postlarvalstage (cyprid) and spend about 2 to 4 wk in the water

    193

    10 m

    20 m 30 m

    38.325N

    38.3215

    38.318

    38.3145

    38.311

    Lat

    itud

    e

    123.084W 123.076 123.068Longitude

    Fig. 1. Map of the study region showing the cross-shelf tran-sect located off Bodega Head in northern California, USA.Circles indicate locations of moorings where plankton towsand CTD casts began. The moorings on the 10, 15, and 22 misobaths included bottom mounted ADCPs and thermistors.The mooring on the 15 m isobath contained a thermistorstring measuring temperature at depths of 4, 7, 10, and 14 m.Wind velocity was measured via an anemometer located on-shore at the Bodega Marine Laboratory, indicated by the

    square

  • Mar Ecol Prog Ser 494: 191203, 2013

    column (Strathmann 1987). Barnacle larval releasegenerally begins in spring (Gaines et al. 1985, Strath-mann 1987) and is continuous through the summermonths (Shanks & Eckert 2005). Barnacle speciesfrom both subtidal (Balanus crenatus, B. nubilus) andintertidal (B. glandula, Chthamalus spp., and Polli -cipes polymerus) habitats are common in this region(Morgan & Fisher 2010). Crab larvae spend weeks tomonths in the plankton, and peak recruitment formost species in this region is during the spring andsummer (Shanks & Eckert 2005, Mace & Morgan2006). The most common taxa include members ofthe families Pinnotheridae, Porcellanidae, Cancri -dae, and Majidae.

    Larval samples

    Cross-shore distributions of nearshore larvae weresampled during 6 daytime cruises using a 0.5 m dia -meter, 200 m mesh net equipped with a mech ani -cal flow meter (Model 2030, General Oceanics). Thenet was modified with a sled to accommodate tow-ing along the bottom. Cruise dates spanned 3 moduring the upwelling season, from May throughJuly 2010, and occurred approximately every 10 d(Fig. 2) under a variety of oceanographic conditions.We sampled 4 stations in a cross-shore transectalong the 10, 15, 22, and 30 m isobaths, correspon-ding to ap proximately 250, 425, 850, and 1100 mfrom shore (Fig. 1). We refer to the 10 and 15 m iso-bath stations as inner CBL stations, and the 22 and30 m isobath stations as outer CBL stations. Weconducted a single 10 min oblique tow at each station, which sampled from the bottom to the sur-face of the water column. Larvae were sorted andidentified to species, or the lowest taxonomic grouppossible, and developmental stage. Larval abun-dances were calculated per m3 to standardize acrossstations.

    Physical data

    To provide physical context during our study, wemeasured currents, temperature, salinity, and winds.Current speed and direction were measured through-out the water column using moored acoustic Dopplercurrent profilers (Workhorse Sentinel ADCP, 1200kHz; Teledyne RD Instruments). Instruments werelocated on the 10, 15, and 22 m isobaths near thestarting position for plankton tows. The ADCPs col-lected 1 min bursts of 0.75 Hz velocity data every 2

    min in 1 m vertical bins that typically extended from~1.5 m above the bottom to ~1.5 m below the surface.The velocity record at the 10 m station ended earlyon 10 June 2012 when its anchor was dislodged. Toquantify general velocity patterns, the raw velocitytime series were depth-averaged, ro tated onto theirprincipal axes, and low-pass filtered with a 33 h cut-off to remove dominant tidal motions (Rosenfeld1983). The major principal axes aligned parallel toshore and along-isobath, corresponding to an angleof 300.

    Bottom temperature was recorded at the ADCPmooring sites every minute over the duration of thestudy at the 10, 15, and 22 m stations, and tempera-tures at depths of 4, 7, 10, and 14 m were recorded atthe 15 m station (SBE 37 and SBE 39, Sea-Bird Elec-tronics). During cruises, temperature, salinity, anddensity were profiled at each station throughout thewater column using a conductivity, temperature, anddepth profiler (SBE 19-Plus, Sea-Bird Electronics),with the exception of the cruise on 25 June.

    Wind data during this study were available from ananemometer located on the shore at the Bodega Mar-ine Laboratory, within 1 km of the study locations, ata height of 20 m (38 19 3.35 N, 123 4 17.20 W;RM Young 05103 Wind Monitor; data available on -line http://bml.ucdavis.edu/boon/).

    Data analysis

    We performed multivariate analyses to determineif larval abundance and larval assemblages variedwith distance from shore and with time. We exam-ined patterns of cross-shore abundance for all taxa,for crab and barnacle larvae separately, and accord-ing to larval stage. All statistical analyses were con-ducted using the multivariate statistical softwarepackage PRIMER v. 6.1.10 (Clarke & Gorley 2006).We determined whether larval assemblages changedwith distance from shore and with sampling dateusing nonparametric analysis of similarity (ANOSIM)and hierarchical cluster analysis and ordination. Datawere fourth-root transformed to reduce the hetero-geneity of variance among samples and assembledinto a Bray- Curtis dissimilarity matrix with a dummyvariable of 1. The resultant dendrogram was testedfor group differences using a similarity profile test(SIMPROF), and the percentage contribution (SIM-PER) of each species and stage to the significant clus-ters was as sessed to classify species-stage combina-tions by their cross-shore distributions and samplingdate. We used non-metric multidimensional scaling

    194

  • Nickols et al.: Larvae in the coastal boundary layer 195

    (NMDS) to exa mine separation of assemblagesaccording to sample date and distance from shore. Toassess if community composition was structured inspace or time, we repeated each of these analyses onuntransformed data that were standardized by totalsample abundance. For all analyses, when there wassignificant structure among samples we then deter-mined which species and stages contributed to thepatterns.

    RESULTS

    Physical conditions

    During the study period, we cap-tured a variety of oceanographic con-ditions, including upwelling, relax-ation, and post-relaxation (Fig. 2).However, our larval sampling datesgenerally occurred during low-windconditions due to logistical constraintson field operations. Daily- averagedalongshore wind speeds during eachsampling date ranged from 8 m s1,indicative of northwesterly winds, to8 m s1, indicative of southeasterlywinds, but winds were predominantlyupwelling favor able over the studyperiod (Fig. 2A). While depth-aver-aged alongshore currents are knownto alternate be tween equatorwardand poleward in this region (Largieret al. 1993, Roughan et al. 2005, Ka-plan & Largier 2006, Morgan et al.2012), on all sampling days of ourstudy the alongshore current waspoleward (Fig. 2B). In general, depth-averaged alongshore velocities meas-ured at inner CBL locations (10 and15 m isobaths) were slower than ve-locities measured at the outer CBL in-strument on the 22 m isobath(Fig. 2B), characteristic of a coastalboundary layer. Exceptions occurredduring onset of strong up wellingwinds on 20 May and 6 June and dur-ing flow reversals.

    Water column properties duringlarval sampling dates ranged fromwell-mixed to stratified (Figs. 2 & 3).The 17 May, 9 June, and 25 Junesampling events represented up well - ing conditions, with a cold homo -geneous water column across sta-

    tions: temperatures below 10C and salinitiessimilarly uniform, with the exception of some low-salinity water on the surface near the outer station on17 May (Figs. 2, 3A & 3C). Alongshore wind speedson the day preceding these sampling dates werefrom the northwest and reached up to 10 m s1, characteristic of strong upwelling conditions andaccounting for the presence of cold isothermal condi-tions over the inner shelf. On 27 May and 4 July, the

    Fig. 2. (A) Alongshore wind velocity measured at Bodega Marine Laboratory.(B) Alongshore depth-averaged and 33 h low-pass filtered current velocitymeasured by bottom-moored ADCPs at the 10, 15, and 22 m isobaths. Positivealong shore velocity is poleward and negative alongshore velocity is equator -ward. (C) Bottom temperature measured on the ADCP moorings at the 10, 15,and 22 m isobaths. (D) Temperature at depths of 4, 7, 10, and 14 m measured at

    the 15 m isobath. Vertical gray bars indicate when sampling occurred

  • Mar Ecol Prog Ser 494: 191203, 2013

    water column was stratified at the 2 inner CBL sta-tions (Figs. 2, 3B & 3D), following a day of southeast-erly or weak winds. After 4 July temperatures in -creased substantially, and the water column wasstratified at all stations on the 14 July sampling day(Fig. 2C). Winds during this period were substantiallyweaker than previous sampling events (Figs. 2C, 2D& 3E).

    Larval abundance

    We identified larvae of 22 crustacean species dur-ing our study (Table 1). The outer CBL station, alongthe 30 m isobath, had the highest number of species,with 10 species on average as compared to 6 to 7 spe-

    cies at the other stations. Throughout the study, lar-vae were most abundant along the 22 m isobath,850 m from shore (Fig. 4A). This pattern was drivenby high abundances of barnacle larvae (Fig. 4B).Early, middle and late stage barnacle larvae werepresent at all stations with the highest abundancealong the 22 m isobath (Fig. 5A). Barnacle postlarvae(cyprids) were found in similar abundance across sta-tions. Crab larvae were most abundant near the 30 misobath, 1100 m from shore (Fig. 4C). All larval crabstages were most abundant at the 2 outer CBL sta-tions, although early stage crab larvae dominated thesamples and abundance decreased with increasingstage (Fig. 5B). Very few crab larvae were found atthe inner CBL stations, and the majority of thosewere early stage larvae.

    196

    0

    10

    20

    30

    0

    10

    20

    30

    0

    10

    20

    30

    0

    10

    20

    30

    0

    10

    20

    30

    Temperature (C) D

    epth

    (m)

    Cross-shore distance (km)

    15

    10

    15

    10

    15

    10

    15

    10

    15

    10

    E

    A

    B

    C

    D

    9

    1011

    9

    9 9

    9

    1011 12

    11 12

    13 14 15

    Salinity

    34

    33

    33.5

    34

    33

    33.5

    34

    33

    33.5

    34

    33

    33.5

    34

    33

    33.5 34

    33.8

    33.8

    34

    34

    33.8 33.6

    33.8

    33.6

    33.4

    0 0.25 0.5 0.75 1 1.25 0 0.25 0.5 0.75 1 1.25

    Fig. 3. Contours of temperature (left) and salinity (right) at the beginning of larval tows along the 10, 15, 22, and 30 m isobathson (A) 17 May 2010, (B) 27 May 2010, (C) 9 June 2010, (D) 4 July 2010, and (E) 14 July 2010. Sample locations are indicated by

    triangles

  • Nickols et al.: Larvae in the coastal boundary layer

    Larval assemblages

    Considering all taxa, larval assemblages within theCBL differed among stations (2-way ANOSIM av =0.322, p = 0.017) and were similar among dates. Be -cause overall larval assemblages were not structuredby date we performed a 1-way ANOSIM to detectspatial differences among assemblages at the differ-ent stations. The innermost station drove differencesbetween assemblages, and it differed from the 2outer stations (1-way ANOSIM pairwise test: for 10 mvs. 30 m stations: R = 0.463, p < 0.01; for 10 m vs. 22 mstations: R = 0.609, p < 0.01). These differences were

    echoed in the dendrogram from cluster analysis, andthe NMDS ordination, which both revealed spatialstructure with 2 main clusters: one defined by lownumbers of larvae, with samples primarily from innerCBL stations, and the second defined by high num-bers of larvae, with samples primarily from the outerCBL stations (Fig. 6).

    Partitioning the analysis by taxa, we found thatpatterns of barnacles and crabs differed. Crab larvalassemblages by themselves did not differ by stationor date, but barnacle larval assemblages did differ bystation and sampling date (2-way ANOSIM: stationav = 0.322, p = 0.02; date av = 0.35, p < 0.01). Thedendrogram from cluster analysis and the NMDSordination revealed 2 main clusters of barnacle sam-ples. One cluster occurred early in the season andmostly offshore, and it was composed of samplesfrom all stations during 17 May and samples from theouter CBL stations during June and July. The othercluster occurred late in the season (June throughJuly) and consisted primarily of samples from theinner CBL stations (Fig. 7). In the early-season, outerCBL assemblage, barnacle larval abundances wereas high as 7200 larvae m3, whereas in the late-season, inner CBL assemblage, barnacle concentra-tions were less than 500 larvae m3.

    We also analyzed changes in the composition of thelarval assemblage. We divided our larval counts (bothspecies and stage) by the total number of larvae ineach sample, providing a fraction of the total samplefor each species and stage, to standardize the data fordifferences in larval abundance across time. Larvalcommunity composition within the CBL differedmostly among stations, but also by date (2-wayANOSIM station av = 0.394, p < 0.01; date av = 0.282,p = 0.026). As noted previously, date was not signifi-cant when abundance was considered in the analysisof larval assemblages. In this analysis of composition,

    197

    10 15 22 300

    1000

    2000

    3000

    Cru

    stac

    ean

    larv

    ae m

    3

    10 15 22 300

    1000

    2000

    3000

    Bar

    nacl

    e la

    rvae

    m3

    10 15 22 300

    100

    200

    300

    Cra

    b la

    rvae

    m3

    Station depth (m)

    CBA

    Fig. 4. Average larval abundance and standard error across date by station for (A) all benthic crustaceans, (B) barnacles, and (C) crabs. Note that the y-axis of (C) is an order of magnitude smaller than (A) and (B)

    Family Taxon

    Cirripedia Balanus crenatus Balanus glandula Balanus nubilus Chthamalus dalli Lepas spp. Pollicipes polymerus

    Cancridae Cancer antennarius Cancer magister Cancer productus Carcinus maenas

    Grapsidae Hemigrapsus oregonensis

    Hippidae Emerita analoga

    Majidae Mimulus foliatus Pugettia producta Pugettia richii Scyra acutifrons

    Paguroidae Pagurid spp.

    Pinnotheridae Pinnotheridae

    Porcellanidae Pachycheles spp. Petrolisthes cinctipes

    Thalassinidae Neotrypaea californiensis

    Xanthidae Lophopanopeus bellus

    Table 1. Crustacean larvae identified in the study

  • Mar Ecol Prog Ser 494: 191203, 2013198

    10 15 22 30 10 15 22 300

    500

    1000

    1500

    2000

    2500B

    arna

    cle

    larv

    ae m

    3

    Station depth (m)

    EarlyMidLatePL

    0

    50

    100

    150

    Cra

    b la

    rvae

    m3

    BA EarlyLatePL

    Fig. 5. Average larval abundance and standard error across date by station for (A) barnacles according to early, mid, late, and postlarval (PL) stage and (B) crabs according to early, late, and PL stage

    Sim

    ilarit

    y (%

    )

    20

    40

    60

    80

    100

    10

    10

    10

    15

    10 10

    15

    15

    30

    15

    22

    15

    30

    22

    30

    22

    30 10

    22 30

    22 15

    30 22

    Similarity 40% 55%

    A

    B15 10 10 10 30 10 15 15 10 30 30 22 30 15 22 22 30 15 10 22 22 15 30 22

    5/17/2010 5/27/2010 6/9/2010 6/25/2010 7/4/2010 7/14/2010

    Fig. 6. (A) Hierarchical clustering dendrogram (using group-average linking) of larval assemblages of benthic crus-taceans from 24 samples taken over 6 d (indicated by sym-bols and legend entry) at 4 sampling stations across thecoastal boundary layer (along the 10, 15, 22, and 30 m iso-baths, indicated by numbers adjacent to symbols), usingtransformed data. Solid black lines indicate significantgroup structure at the 5% level. Dashed lines represent non-significant group structure. Sample station isobaths are re-ported beneath each symbol. (B) Non-metric multidimen-sional scaling plot (2D stress, 0.15; 3D stress, 0.09) from the24 samples with superimposed significant clusters at simi -larity levels of 40% (solid lines) and 55% (dashed lines).

    Symbols and numbers as in (A)

    22 30 15 30 10 10 30 10 15 15 10 30 30 22 22 15 10 22 22 15 10 22 15 30

    Sim

    ilarit

    y (%

    )

    20

    40

    60

    80

    100

    5/17/2010 5/27/2010 6/9/2010 6/25/2010 7/4/2010 7/14/2010

    A

    B

    10

    10 10

    15

    15

    15 10

    10

    10 15

    15

    15

    22

    22

    22

    22

    22 22

    30

    30

    30 30

    30

    30

    Similarity 55% 75%

    Fig. 7. (A) Hierarchical clustering dendrogram (using group-average linking) of barnacle larval assemblages from 24samples taken over 6 d (indicated by symbols and legendentry) at 4 sampling stations across the coastal boundarylayer (along the 10, 15, 22, and 30 m isobaths, indicated bynumbers adjacent to symbols). Solid black lines indicate sig-nificant group structure at the 5% level. Dashed lines repre-sent non-significant group structure. Sample station iso-baths are reported beneath each symbol. (B) Non-metricmultidimensional scaling plot (2D stress, 0.12; 3D stress,0.07) from the 24 samples with superimposed significantclusters at simi larity levels of 55% (solid lines) and 75%

    (dashed lines). Symbols and numbers as in (A)

  • Nickols et al.: Larvae in the coastal boundary layer

    the effect of date was driven by one particular sam-pling event on 9 June, when there was a distinct innerCBL assemblage (10 and 15 m stations; 92% similar-ity). The dendrogram from cluster analysis and theNMDS ordination showed 2 additional as semblages(Fig. 8): an inner CBL assemblage (68% similarity)and a main CBL assemblage containing nearly all ofthe other samples from the CBL, the 2 outer CBL sta-tions as well as the 15 m station (67% similarity).

    The barnacles Balanus crenatus and B. nubilus lar-vae dominated the composition of the assemblageand drove differences between clusters. The assem-blage at the inner CBL stations on 9 June (Fig. 8) wascomposed of >75% B. crenatus cyprids. All other

    samples were composed of 30% or less B. crenatuscyprids. In addition, the 9 June inner CBL assem-blage contained

  • Mar Ecol Prog Ser 494: 191203, 2013200

    da (2007) measured larval concentrations of Balanusglandula and Chtha ma lus spp. at 3 cross-shore sta-tions within 1100 m from shore over a period of 7 d.While concentrations of most larval stages of B. glan-dula were similar among stations, concentrations ofthird through sixth stage Chthamalus nauplii werelower at the innermost station, 300 m from shore(Tapia & Pineda 2007). Even over a short temporalperiod, barnacle concentrations exhibited spatialstructure and nauplii may potentially have avoidedvery nearshore waters.

    Despite low concentrations of larvae in the inner-most waters of the CBL, the high concentrations of alllarval stages of barnacles in the CBL along the 15, 22,and 30 m isobaths suggest that many barnacle larvaemay be retained within the CBL and develop in waters within 1100 m from shore. All larval stagesof crabs occurred at the 2 outer CBL stations, sug-gesting that they may also complete developmentwithin the CBL. These findings are consistent withother studies in both weak and strong upwelling re-gions where high abundances of all larval stages oc-curred within a few kilometers from shore (Tapia &Pineda 2007, Morgan et al. 2009b,c, Shanks & Shear-man 2009, Morgan & Fisher 2010, Morgan et al. 2011,2012). Benthic crustacean larvae exhibit depth pref-erences that can aid nearshore retention for mostspecies in upwelling regions (Miller & Morgan 2013).By remaining near the bottom, larvae can take ad-vantage of slower velocities in the bottom boundarylayer (in both the along- and cross-shore directions),as well as avoid offshore transport in the surface Ek-man layer (Morgan et al. 2009b, Shanks & Shearman2009, Morgan & Fisher 2010, Morgan et al. 2012).

    Inshore and offshore assemblages within the CBL

    A distinct larval assemblage occurred closest toshore within the CBL. Although some features of lar-val assemblages changed with time, these assem-blages were predominantly defined by space: notonly was there spatial structure in larval abundance,there was spatial structure in the composition of theas semblage. This spatial structure occurred eventhough physical conditions were variable amongsampling dates (flow velocity, water temperature,stratification) and on many days there was no cleardifference in physical parameters between the inner-most station and those within the rest of the CBL(Figs. 2 & 3).

    The spatial boundary between assemblages ofinner and outer stations within the CBL is dynamic.

    Larval assemblages on half of the sampling dates atthe 15 m station were most similar to the 10 m station,and on the other half were more similar to the 22 mstation. There is no clear physical difference betweenthese groupings of days apparent from our data, asthey spanned oceanographic conditions. For exam-ple, on 9 June, 4 July, and 14 July the 15 m stationmatched most closely with the 10 m station, yet thewater column profiles from each of these days arequite distinct (Fig. 3CE). One possible physical fac-tor we did not explore that could influence thedemarcation of the inshore community is the width ofthe surf zone and associated rip current zone, whichis itself a dynamic boundary, dependent on the sig-nificant wave height and tidal elevation (Lentz et al.1999, Brown et al. 2009). Surf zone characteristicsappear to impact shoreline settlement of inverte-brates, with low settlement observed at reflectivebeaches which are characterized by high beachslopes and standing waves and are thought to havereduced cross-shore exchange (Shanks et al. 2010).Rocky shores are hypothesized to be similar to reflec-tive beaches, and if so, the associated reduction incross-shore ex change might explain low settlementat some locations and low abundances of larvae insurf zone waters (Shanks et al. 2010). Our studyfound low larval concentrations in waters just beyondthe surf zone, but at distances that could be influ-enced by surf zone processes through the action ofrip currents. Specifically, off Horseshoe Cove (Fig. 1,just downcoast of station locations), wave-driven cir-culation has been observed to extend as a macro-ripup to distances ~250 m offshore (J. L. Largier,unpubl. drifter data), comparable with the distance toour inner CBL station along the 10 m isobath. Thissuggests that the influence of wave-driven processeson larval transport may extend offshore (contrary tothe idea of reduced exchange off rocky shores, assuggested by Shanks et al. 2010).

    In addition to potential physical differences be -tween the habitat of the inshore and offshore as -semblages, predation may be higher within the nar-row band of inner CBL water than farther offshore.Habitat along the 10 m isobath at our study site fea-tures rocky substrate with some areas supportingstands of the bull kelp, Nereocystis luet keana. Incentral California, larval abundances were found tobe negatively correlated with kelp density, and lowerlarval abundances on the inshore edges of kelpforests were attributed to predation (Gaines &Rough garden 1987). Although the kelp in our regionis much more sparse than the giant kelp Macrocystispyrifera beds in central California, predation is still a

  • Nickols et al.: Larvae in the coastal boundary layer

    possible explanation for decreased abundance at themost inshore station of our study.

    Implications of cross-shelf larval structurewithin the CBL

    Larvae are clearly spending time within the coastalboundary layer; some may even complete their entiredevelopment within the CBL, which could impactestimates of population connectivity. During theirtime in the CBL, larvae are exposed to slower movingalongshore flows than farther offshore, which willhave an impact on overall dispersal distance (Nickolset al. 2012, Nickols et al. unpubl. data). Although wedid not have current velocity measurements through-out the water column beyond the 22 m isobath, con-current measurements of surface currents by high- frequency radar showed that current velocities werefaster farther offshore (data not shown). The radardomain begins 2 km offshore, and generally has highagreement with measurements from ADCPs (Kaplanet al. 2005). This gradient is also observed in a cross-shore array of moorings deployed during WEST(Wind Events and Shelf Transport; Largier et al.2006) and in other unpublished data from BML. Esti-mates of dispersal distance in this region shouldtherefore consider current velocities within 1 km orless from shore, as this is where the majority of larvaeappear to be concentrated. Such consideration mayimprove estimates of dispersal distance derived frompelagic larval durations, which are often larger thandispersal distances estimated from genetics, tagging,and natural tracers (Palumbi 2004, Jones et al. 2009,Shanks 2009, Lpez-Duarte et al. 2012). Refining ourunderstanding of dispersal distances will improveour ability to accurately model population dynamicsand assess population persistence (Botsford et al.2009, White et al. 2010, Burgess et al. in press).

    The coast of northern California generally haslower recruitment than other regions along the westcoast of North America (Connolly et al. 2001), and alongstanding question has been whether or not thispattern is linked to larval supply. Although it wasproposed that larval supply is diminished when lar-vae are forced offshore by strong upwelling (e.g.Roughgarden et al. 1988), numerous studies nowsuggest strongly that many larvae of multiple speciesare retained nearshore during both upwelling andrelaxation conditions (Tapia & Pineda 2007, Morganet al. 2009b,c, Shanks & Shearman 2009, Morgan &Fisher 2010, Morgan et al. 2011). Our study alsofound high abundance of larvae close to the shore,

    and extended closer to shore than previous work inthis region of strong upwelling.

    An important finding of our study is the observa-tion of low larval concentrations and a different lar-val assemblage in the innermost waters of the CBL,indicating a potential disconnect between high larvalabundance in the CBL and larval supply to shorelinerecruitment habitat. Further, this disconnect appearsto occur in waters beyond the surf zone, in contrast torecent work by Shanks et al. (2010) that suggests thatsurf zone processes may disrupt the supply of near -shore planktonic larvae to shoreline habitats. Whilethese results are from a single location, they repre-sent a diversity of oceanographic conditions and theobserved mismatch raises important questions abouthow general this result may be. Our study also fo -cuses attention on the need to understand the mech-anisms that control transport of larvae to shorelinehabitats, while highlighting methodological concernsof studies that explore links between supply and set-tlement. As we endeavor to better understand thelinks between larval dispersal and population dy -namics, it is essential that the nearshore zone bestudied in greater detail and that we work to addressthe spatial pattern of recruitment limitation in coastalsystems.

    Acknowledgements. We thank J. Demmer, R. Fontana, andN. Weidberg for assistance in the field as well as D. Dann,M. Robart, and J. Herum for help with deployments ofoceanographic instrumentation. J. Fisher provided guidanceand feedback. This work was funded by California SeaGrant (NA08AR4170669) and the National Science Founda-tion (OCE-0927196 and OCE-0927255). K.J.N. was also sup-ported by a Bodega Marine Laboratory Graduate StudentFellowship.

    LITERATURE CITED

    Bertness MD, Gaines SD, Stephens EG, Yund PO (1992)Components of recruitment in populations of the acornbarnacle Semibalanus balanoides (Linnaeus). J Exp MarBiol Ecol 156: 199215

    Borges R, Ben-Hamadou R, Chcharo MA, R P, GonalvesEJ (2007) Horizontal spatial and temporal distributionpatterns of nearshore larval fish assemblages at a tem-perate rocky shore. Estuar Coast Shelf Sci 71: 412428

    Botsford LW (2001) Physical influences on recruitment toCalifornia Current invertebrate populations on multiplescales. ICES J Mar Sci 58: 10811091

    Botsford LW, Moloney CL, Hastings A, Largier JL, PowellTM, Higgins K, Quinn JF (1994) The influence of spa-tially and temporally varying oceanographic conditionson meroplankton metapopulations. Deep-Sea Res II 41: 107145

    Botsford LW, White JW, Coffroth MA, Paris CB and others(2009) Connectivity and resilience of coral reef metapop-

    201

    http://dx.doi.org/10.1007/s00338-009-0466-zhttp://dx.doi.org/10.1016/0967-0645(94)90064-7http://dx.doi.org/10.1006/jmsc.2001.1085http://dx.doi.org/10.1016/j.ecss.2006.08.020http://dx.doi.org/10.1016/0022-0981(92)90246-7

  • Mar Ecol Prog Ser 494: 191203, 2013

    ulations in marine protected areas: matching empiricalefforts to predictive needs. Coral Reefs 28: 327337

    Brown J, MacMahan J, Reniers A, Thornton E (2009) Surfzone diffusivity on a rip-channeled beach. J Geophys Res114: C11015, doi: 10.1029/2008JC005158

    Burgess SC, Nickols KJ, Griesemer CD, Barnett LAK andothers (in press) Beyond connectivity: how empiricalmethods can quantify population persistence to improvemarine protected area design. Ecol Appl

    Clarke KR, Gorley RN (2006) PRIMER v6: User Manual/Tutorial. PRIMER-E, Plymouth

    Connell JH (1985) The consequences of variation in initialsettlement vs. post-settlement mortality in rocky inter-tidal communities. J Exp Mar Biol Ecol 93: 1145

    Connolly SR, Menge BA, Roughgarden J (2001) A latitudinalgradient in recruitment of intertidal invertebrates in thenortheast Pacific Ocean. Ecology 82: 17991813

    Cowen RK (1985) Large-scale pattern of recruitment by thelabrid, Semicossyphus pulcher : causes and implications.J Mar Res 43: 719742

    Dudas SE, Rilov G, Tyburczy J, Menge BA (2009) Linkinglarval abundance, onshore supply and settlement usinginstantaneous versus integrated methods. Mar Ecol ProgSer 387: 8195

    Farrell TM, Bracher D, Roughgarden J (1991) Cross-shelftransport causes recruitment to intertidal populations incentral California. Limnol Oceanogr 36: 279288

    Gaines SD, Bertness MD (1992) Dispersal of juveniles andvariable recruitment in sessile marine species. Nature360: 579580

    Gaines SD, Roughgarden J (1985) Larval settlement rate: aleading determinant of structure in an ecological com-munity of the marine intertidal zone. Proc Natl Acad SciUSA 82: 37073711

    Gaines SD, Roughgarden J (1987) Fish in offshore kelpforests affect recruitment to intertidal barnacle popula-tions. Science 235: 479481

    Gaines SD, Brown S, Roughgarden J (1985) Spatial variationin larval concentrations as a cause of spatial variation insettlement for the barnacle, Balanus glandula. Oecologia67: 267272

    Gaylord B, Gaines SD (2000) Temperature or transport?Range limits in marine species mediated solely by flow.Am Nat 155: 769789

    Gaylord B, Reed DC, Raimondi PT, Washburn L, McLean SR(2002) A physically based model of macroalgal spore dis-persal in the wave and current-dominated nearshore.Ecology 83: 12391251

    Gaylord B, Reed DC, Washburn L, Raimondi PT (2004) Phys-ical-biological coupling in spore dispersal of kelp forestmacroalage. J Mar Syst 49: 1939

    Gaylord B, Reed DC, Raimondi PT, Wasburn L (2006)Macroalgal spore dispersal in coastal environments: mechanistic insights revealed by theory and experiment.Ecol Monogr 76: 481502

    Gaylord B, Rosman JH, Reed DC, Koseff JR and others(2007) Spatial patterns of flow and their modificationwithin and around a giant kelp forest. Limnol Oceanogr52: 18381852

    Gaylord B, Nickols KJ, Jurgens L (2012) Roles of transportand mixing processes in kelp forest ecology. J Exp Biol215: 9971007

    Graham WM, Largier JL (1997) Upwelling shadows as near -shore retention sites: the example of northern MontereyBay. Cont Shelf Res 17: 509532

    Jones GP, Almany GR, Russ GR, Sale PF, Steneck RS, vanOppen MJH, Willis BL (2009) Larval retention and con-nectivity among populations of corals and reef fishes: his-tory, advances and challenges. Coral Reefs 28: 307325

    Kaplan DM, Largier JL (2006) HF radar-derived origin andsurface waters off Bodega Bay, California. Deep-Sea ResII 53: 29062930

    Kaplan DM, Largier JL, Botsford LW (2005) HF radar obser-vations of surface circulation off Bodega Bay (northernCalifornia, USA). J Geophys Res 110: C10020, doi: 10.1029/ 2005JC002959

    Largier JL (2003) Considerations in estimating larval disper-sal distances from oceanographic data. Ecol Appl 13: S71S89

    Largier JL, Magnell BA, Winant CD (1993) Subtidal circula-tion over the Northern California shelf. J Geophys Res98: 1814718179

    Largier JL, Lawrence CA, Roughan M, Kaplan DM and oth-ers (2006) WEST: a northern Califrnia study of the role ofwind-driven transport in the productivity of coastalplankton communities. Deep-Sea Res II 53: 28332849

    Lentz SJ, Guza RT, Elgar S, Feddersen F, Herbers THC(1999) Momentum balances on the North Carolina innershelf. J Geophys Res 104: 1820518226

    Levin LA (2006) Recent progress in understanding larvaldispersal: new directions and digressions. Integr CompBiol 46: 282297

    Lpez-Duarte PC, Carson HS, Cook GS, Fodrie FJ, BeckerBJ, DiBacco C, Levin LA (2012) What controls connectiv-ity? An empirical, multi-species approach. Integr CompBiol 52: 511524

    Mace AJ, Morgan SG (2006) Larval accumulation in the leeof a small headland: implications for the design of marinereserves. Mar Ecol Prog Ser 318: 1929

    MacMahan J, Brown J, Brown J, Thornton E and others(2010) Mean Lagrangian flow behavior on an open coastrip-channeled beach: a new perspective. Mar Geol 268: 115

    McCulloch A, Shanks AL (2003) Topographically generatedfronts, very nearshore oceanography and the distributionand settlement of mussel larvae and barnacle cyprids.J Plankton Res 25: 14271439

    McPhee-Shaw EE, Nielsen KJ, Largier JL, Menge BA (2011)Nearshore chlorophyll-a events and wave-driven trans-port. Geophys Res Lett 38: 15

    McQuaid CD, Phillips TE (2000) Limited wind-driven dis-persal of intertidal mussel larvae: in situ evidence fromthe plankton and the spread of the invasive speciesMytilus galloprovincialis in South Africa. Mar Ecol ProgSer 201: 211220

    Menge BA, Blanchette C, Raimondi P, Freidenburg T andothers (2004) Species interaction strength: testing modelpredictions along an upwelling gradient. Ecol Monogr74: 663684

    Miller SH, Morgan SG (2013) Interspecific differences indepth preference: regulation of larval transport in anupwelling system. Mar Ecol Prog Ser 476: 301306

    Monismith SG (2004) A note on the potential transport ofscalars and organisms by surface waters. LimnolOceanogr 49: 12141217

    Morgan SG (2001) The larval ecology of marine communi-ties. In: Bertness MD, Gaines SD, Hay ME (eds) Marinecommunity ecology. Sinauer, Sunderland, MA

    Morgan SG, Fisher JL (2010) Larval behavior regulatesnearshore retention and offshore migration in an up -

    202

    http://dx.doi.org/10.3354/meps08476http://dx.doi.org/10.4319/lo.2004.49.4.1214http://dx.doi.org/10.3354/meps10150http://dx.doi.org/10.1890/03-4060http://dx.doi.org/10.3354/meps201211http://dx.doi.org/10.1029/2010GL045810http://dx.doi.org/10.1093/plankt/fbg098http://dx.doi.org/10.1016/j.margeo.2009.09.011http://dx.doi.org/10.3354/meps318019http://dx.doi.org/10.1093/icb/ics104http://dx.doi.org/10.1093/icb/icj024http://dx.doi.org/10.1029/1999JC900101http://dx.doi.org/10.1016/j.dsr2.2006.08.018http://dx.doi.org/10.1029/93JC01074http://dx.doi.org/10.1890/1051-0761(2003)013[0071%3ACIELDD]2.0.CO%3B2http://dx.doi.org/10.1029/2005JC002959http://dx.doi.org/10.1016/j.dsr2.2006.07.012http://dx.doi.org/10.1007/s00338-009-0469-9http://dx.doi.org/10.1016/S0278-4343(96)00045-3http://dx.doi.org/10.1242/jeb.059824http://dx.doi.org/10.4319/lo.2007.52.5.1838http://dx.doi.org/10.1890/0012-9615(2006)076[0481%3AMSDICE]2.0.CO%3B2http://dx.doi.org/10.1016/j.jmarsys.2003.05.003http://dx.doi.org/10.1890/0012-9658(2002)083[1239%3AAPBMOM]2.0.CO%3B2http://dx.doi.org/10.1086/303357http://dx.doi.org/10.1007/BF00384297http://dx.doi.org/10.1126/science.235.4787.479http://dx.doi.org/10.1073/pnas.82.11.3707http://dx.doi.org/10.1038/360579a0http://dx.doi.org/10.4319/lo.1991.36.2.0279http://dx.doi.org/10.3354/meps08095http://dx.doi.org/10.1357/002224085788440376http://dx.doi.org/10.1890/0012-9658(2001)082[1799%3AALGIRO]2.0.CO%3B2http://dx.doi.org/10.1016/0022-0981(85)90146-7http://dx.doi.org/10.1029/2008JC005158

  • Nickols et al.: Larvae in the coastal boundary layer

    welling shadow and along the open coast. Mar Ecol ProgSer 404: 109126

    Morgan SG, Fisher JL, Mace AJ (2009a) Larval recruitmentin a region of strong, persistent upwelling and recruit-ment limitation. Mar Ecol Prog Ser 394: 7999

    Morgan SG, Fisher JL, Miller SH, McAfee ST, Largier JL(2009b) Nearshore larval retention in a region of strongupwelling and recruitment limitation. Ecology 90: 34893502

    Morgan SG, Fisher JL, Mace AJ, Akins L, Slaughter AM,Bollens SM (2009c) Cross-shelf distributions and recruit-ment of crab postlarvae in a region of strong upwelling.Mar Ecol Prog Ser 380: 173185

    Morgan SG, Fisher JL, Largier JL (2011) Larval retention,entrainment and accumulation in the lee of a small head-land: recruitment hotspots along windy coasts. LimnolOceanogr 56: 161178

    Morgan SG, Fisher JL, McAfee ST, Largier JL, Halle CM(2012) Limited recruitment during relaxation events: lar-val advection and behavior in an upwelling system. Lim-nol Oceanogr 57: 457470

    Nickols KJ, Gaylord B, Largier JL (2012) The coastal bound-ary layer: predictable current structure decreases along-shore transport and alters scales of dispersal. Mar EcolProg Ser 464: 1735

    Palumbi SR (2004) Marine reserves and ocean neighbor-hoods: the spatial scale of marine populations and theirmanagement. Annu Rev Environ Resour 29: 3168

    Prairie JC, Sutherland KR, Nickols KJ, Kaltenberg AM(2012) Biophysical interactions in the plankton: a cross-scale review. Limnol Oceanogr Fluids Environ 2: 121145

    Reed DC, Laur DR, Ebeling AW (1988) Variation in algal dis-persal and recruitment: the importance of episodicevents. Ecol Monogr 58: 321335

    Reed DC, Kinlan BP, Raimondi PT, Washburn L, Gaylord B,Drake PT (2006) A metapopulation perspective on patchdynamics and connectivity of giant kelp. In: Kritzer JP,Sale PF (eds) Marine metapopulations. Academic Press,San Diego, CA

    Rilov G, Dudas SE, Menge BA, Grantham BA, Lubchenco J,Schiel DR (2008) The surf zone: a semi-permeable bar-rier to onshore recruitment of invertebrate larvae? J ExpMar Biol Ecol 361: 5974

    Rosenfeld LK (1983) CODE-2: Moored array and large-scaledata report. WHOI Technical Report 85-35: 1242

    Roughan M, Mace AJ, Largier JL, Morgan SG, Fisher JL,Carter ML (2005) Subsurface recirculation and larvalretention in the lee of a small headland: a variation on theupwelling shadow theme. J Geophys Res 110: C10027,doi: 10.1029/2005JC002898

    Roughgarden J, Gaines SD, Possingham H (1988) Recruit-ment dynamics in complex life-cycles. Science 241: 14601466

    Send U, Beardsley RC, Winant CD (1987) Relaxation fromupwelling in the Coastal Ocean Dynamics Experiment.J Geophys Res 92: 16831698

    Shanks AL (2009) Pelagic larval duration and dispersal dis-

    tance revisited. Biol Bull 216: 373385Shanks AL, Eckert GL (2005) Population persistence of Cal-

    ifornia Current fishes and benthic crustaceans: a marinedrift paradox. Ecol Monogr 75: 505524

    Shanks AL, Shearman RK (2009) Paradigm lost? Cross-shelfdistributions of intertidal invertebrate larvae are unaf-fected by upwelling or downwelling. Mar Ecol Prog Ser385: 189204

    Shanks AL, Morgan SG, MacMahan J, Reniers AJHM(2010) Surf zone physical and morphological regime asdeterminants of temporal and spatial variation in larvalrecruitment. J Exp Mar Biol Ecol 392: 140150

    Strathmann MF (1987) Reproduction and development ofmarine invertebrates of the northern Pacific Coast. Uni-versity of Washington Press, Seattle, WA

    Swearer SE, Shima JS, Hellberg ME, Thorrold SR and others(2002) Evidence of self-recruitment in demersal marinepopulations. Bull Mar Sci 70: S251S271

    Tapia FJ, Pineda J (2007) Stage-specific distribution of bar-nacle larvae in nearshore waters: potential for limiteddispersal and high mortality rates. Mar Ecol Prog Ser342: 177190

    Thorson G (1946) Reproduction and larval development ofDanish marine bottom invertebrates. Meddelelser fraKommissionen for Hovundersoegelser Serie Plankton 4: 1523

    Underwood AJ, Fairweather PG (1989) Supply-side ecologyand benthic marine assemblages. Trends Ecol Evol 4: 1620

    Underwood AJ, Keough MJ (2001) Supply-side ecology: thenature and consequences of variations in recruitment ofintertidal organisms. In: Bertness MD, Gaines SD, HayME (eds) Marine community ecology. Sinauer, Sunder-land, MA

    White JW, Botsford LW, Hastings A, Largier JL (2010) Popu-lation persistence in marine reserve networks: incorpo-rating spatial heterogeneities in larval dispersal. MarEcol Prog Ser 398: 4967

    Winant CD, Beardsley RC, Davis RE (1987) Moored wind,temperature, and current observations made duringCoastal Ocean Dynamics Experiments 1 and 2 over theNorthern California continental shelf and upper slope.J Geophys Res 92: 15691604

    Wing SR, Botsford LW, Largier JL, Morgan LE (1995a) Spa-tial structure of relaxation events and crab settlement inthe northern California upwelling system. Mar Ecol ProgSer 128: 199211

    Wing SR, Largier JL, Botsford LW, Quinn JF (1995b) Settle-ment and transport of benthic invertebrates in an intermittent upwelling region. Limnol Oceanogr 40: 316329

    Wing SR, Botsford LW, Ralston SV, Largier JL (1998) Mero-planktonic distribution and circulation in a coastal reten-tion zone of the northern California upwelling system.Limnol Oceanogr 43: 17101721

    Yoshioka PM (1982) Role of planktonic and benthic factorsin the population dynamics of the bryozoan Membrani-pora membranacea. Ecology 63: 457468

    203

    Editorial responsibility: Lisandro Benedetti-Cecchi, Pisa, Italy

    Submitted: March 28, 2013; Accepted: September 20, 2013Proofs received from author(s): November 14, 2013

    http://dx.doi.org/10.2307/1938963http://dx.doi.org/10.4319/lo.1998.43.7.1710http://dx.doi.org/10.4319/lo.1995.40.2.0316http://dx.doi.org/10.3354/meps128199http://dx.doi.org/10.1029/JC092iC02p01569http://dx.doi.org/10.3354/meps08327http://dx.doi.org/10.1016/0169-5347(89)90008-6http://dx.doi.org/10.3354/meps342177http://dx.doi.org/10.1016/j.jembe.2010.04.018http://dx.doi.org/10.3354/meps08043http://dx.doi.org/10.1890/05-0309http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=19556601&dopt=Abstracthttp://dx.doi.org/10.1029/JC092iC02p01683http://dx.doi.org/10.1126/science.11538249http://dx.doi.org/10.1016/j.jembe.2008.04.008http://dx.doi.org/10.2307/1942543http://dx.doi.org/10.1215/21573689-1964713http://dx.doi.org/10.1146/annurev.energy.29.062403.102254http://dx.doi.org/10.3354/meps09875http://dx.doi.org/10.4319/lo.2012.57.2.0457http://dx.doi.org/10.4319/lo.2011.56.1.0161http://dx.doi.org/10.3354/meps07913http://dx.doi.org/10.1890/08-1550.1http://dx.doi.org/10.3354/meps08216

    cite10: cite12: cite21: cite23: cite30: cite18: cite32: cite27: cite41: cite4: cite43: cite36: cite8: cite38: cite52: cite50: cite54: cite34: cite47: cite56: cite61: cite45: cite58: cite29: cite63: cite65: cite70: cite72: cite69: cite74: cite1: cite9: cite11: cite13: cite20: cite22: cite15: cite24: cite17: cite2: cite26: cite40: cite33: cite28: cite42: cite6: cite51: cite31: cite44: cite39: cite53: cite46: cite55: cite48: cite60: cite57: cite62: cite64: cite66: cite71: cite73: cite68: cite3: cite7: