16
RESEARCH ARTICLE 10.1002/2014JC010464 Field observations of wave-driven circulation over spur and groove formations on a coral reef Justin S. Rogers 1 , Stephen G. Monismith 1 , Robert B. Dunbar 2 , and David Koweek 2 1 Environmental Fluid Mechanics Laboratory, Stanford University, Stanford, California, USA, 2 Department of Environmental Earth System Science, Stanford University, Stanford, California, USA Abstract Spur and groove (SAG) formations are found on the forereefs of many coral reefs worldwide. Modeling results have shown that SAG formations together with shoaling waves induce a nearshore Lagran- gian circulation pattern of counter-rotating circulation cells, but these have never been observed in the field. We present results from two separate field studies of SAG formations on Palmyra Atoll which show their effect on waves to be small, but reveal a persistent order 1 cm/s depth-averaged Lagrangian offshore flow over the spur and onshore flow over the grooves. This circulation was stronger for larger, directly inci- dent waves and low alongshore flow conditions, consistent with predictions from modeling. Favorable forc- ing conditions must be maintained on the order of 1 h to accelerate and develop the SAG circulation cells. The primary cross and alongshore depth-averaged momentum balances were between the pressure gradi- ent, radiation stress gradient, and nonlinear convective terms, and the bottom drag was similar to values found on other reefs. The vertical structure of these circulation cells was previously unknown and the results show a complex horizontal offshore Lagrangian flow over the spurs near the surface driven by alongshore variability in radiation stress gradients. Vertical flow was downward over the spur and upward over the groove, likely driven by alongshore differences in bottom stress and not by vortex forcing. 1. Introduction One of the most prominent features of coral reefs worldwide are elevated periodic shore-normal ridges of coral (‘‘spurs’’) separated by shore-normal patches of sediment (‘‘grooves’’), generally located on the forereef offshore of the surf zone [Storlazzi et al., 2003]. These features, termed ‘‘spur and groove’’ (SAG) formations, have been observed on fringing reefs, barrier reefs, and atolls, and vary in their size and shape [Rogers et al., 2013]. Hydrodynamic processes influence coral growth in several ways [Chappell, 1980]. First, waves and mean flows can suspend and transport sediments and reef debris. Suspended sediment is generally recognized as an important factor that can negatively affect coral health [Buddemeier and Hopley, 1988; Acevedo et al., 1989; Rogers, 1990; Fortes, 2000; Fabricius, 2005]. Second, forces imposed by waves can subject corals to breakage, resulting in trimming or reconfiguration of the reef [Masselink and Hughes, 2003; Storlazzi et al., 2005]. Third, the rates of nutrient uptake on coral reefs [Atkinson and Bilger, 1992; Thomas and Atkinson, 1997; Atkinson et al., 2001], photosynthetic production, and calcification by coral [Dennison and Barnes, 1988] and particulate capture by reef organisms [Genin et al., 2009] increase with increasing water motion. Although the geometric properties of SAG formations are well documented [Munk and Sargent, 1954; Rob- erts, 1974; Blanchon and Jones, 1997; Storlazzi et al., 2003], analysis of their hydrodynamic function has been limited [Rogers et al., 2013]. Using a depth-averaged, phase-resolving model (funwaveC), Rogers et al. [2013] showed that SAG formations together with shoaling waves induce a nearshore Lagrangian circulation pat- tern of counter-rotating circulation cells, confirming suggestions from the geologic literature [Munk and Sar- gent, 1954; Roberts et al., 1977; Storlazzi et al., 2003]. This model also revealed that this type of circulation is enhanced by spur-normal waves, increased wave height, weak alongshore currents, increased spur height, and decreased bottom drag. The classical dynamical basis by which waves drive flow is through changes to the waves from physical processes such as shoaling, refraction, dissipation, etc., which create spatial gradients in radiation stresses and impart a force in the momentum equation [Longuet-Higgins and Stewart, 1964]. Specifically for SAG Key Points: Waves over spur and groove formations induce Lagrangian circulation cells Horizontal flow is offshore over spurs and weak onshore over grooves Downward flow over spur is likely due to variation in alongshore bottom stress Correspondence to: J. S. Rogers, [email protected] Citation: Rogers, J. S., S. G. Monismith, R. B. Dunbar, and D. Koweek (2015), Field observations of wave-driven circulation over spur and groove formations on a coral reef, J. Geophys. Res. Oceans, 120, doi:10.1002/ 2014JC010464. Received 19 SEP 2014 Accepted 8 DEC 2014 Accepted article online 16 DEC 2014 ROGERS ET AL. V C 2014. American Geophysical Union. All Rights Reserved. 1 Journal of Geophysical Research: Oceans PUBLICATIONS

Field observations of wave‐driven circulation over spur ...web.stanford.edu/~jsrogers/rogers et al 2015 - field observations of wave-driven...Abstract Spur and groove (SAG) formations

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

RESEARCH ARTICLE10.1002/2014JC010464

Field observations of wave-driven circulation over spur andgroove formations on a coral reefJustin S. Rogers1, Stephen G. Monismith1, Robert B. Dunbar2, and David Koweek2

1Environmental Fluid Mechanics Laboratory, Stanford University, Stanford, California, USA, 2Department of EnvironmentalEarth System Science, Stanford University, Stanford, California, USA

Abstract Spur and groove (SAG) formations are found on the forereefs of many coral reefs worldwide.Modeling results have shown that SAG formations together with shoaling waves induce a nearshore Lagran-gian circulation pattern of counter-rotating circulation cells, but these have never been observed in thefield. We present results from two separate field studies of SAG formations on Palmyra Atoll which showtheir effect on waves to be small, but reveal a persistent order 1 cm/s depth-averaged Lagrangian offshoreflow over the spur and onshore flow over the grooves. This circulation was stronger for larger, directly inci-dent waves and low alongshore flow conditions, consistent with predictions from modeling. Favorable forc-ing conditions must be maintained on the order of 1 h to accelerate and develop the SAG circulation cells.The primary cross and alongshore depth-averaged momentum balances were between the pressure gradi-ent, radiation stress gradient, and nonlinear convective terms, and the bottom drag was similar to valuesfound on other reefs. The vertical structure of these circulation cells was previously unknown and the resultsshow a complex horizontal offshore Lagrangian flow over the spurs near the surface driven by alongshorevariability in radiation stress gradients. Vertical flow was downward over the spur and upward over thegroove, likely driven by alongshore differences in bottom stress and not by vortex forcing.

1. Introduction

One of the most prominent features of coral reefs worldwide are elevated periodic shore-normal ridges ofcoral (‘‘spurs’’) separated by shore-normal patches of sediment (‘‘grooves’’), generally located on the forereefoffshore of the surf zone [Storlazzi et al., 2003]. These features, termed ‘‘spur and groove’’ (SAG) formations,have been observed on fringing reefs, barrier reefs, and atolls, and vary in their size and shape [Rogers et al.,2013].

Hydrodynamic processes influence coral growth in several ways [Chappell, 1980]. First, waves and meanflows can suspend and transport sediments and reef debris. Suspended sediment is generally recognized asan important factor that can negatively affect coral health [Buddemeier and Hopley, 1988; Acevedo et al.,1989; Rogers, 1990; Fortes, 2000; Fabricius, 2005]. Second, forces imposed by waves can subject corals tobreakage, resulting in trimming or reconfiguration of the reef [Masselink and Hughes, 2003; Storlazzi et al.,2005]. Third, the rates of nutrient uptake on coral reefs [Atkinson and Bilger, 1992; Thomas and Atkinson,1997; Atkinson et al., 2001], photosynthetic production, and calcification by coral [Dennison and Barnes,1988] and particulate capture by reef organisms [Genin et al., 2009] increase with increasing water motion.

Although the geometric properties of SAG formations are well documented [Munk and Sargent, 1954; Rob-erts, 1974; Blanchon and Jones, 1997; Storlazzi et al., 2003], analysis of their hydrodynamic function has beenlimited [Rogers et al., 2013]. Using a depth-averaged, phase-resolving model (funwaveC), Rogers et al. [2013]showed that SAG formations together with shoaling waves induce a nearshore Lagrangian circulation pat-tern of counter-rotating circulation cells, confirming suggestions from the geologic literature [Munk and Sar-gent, 1954; Roberts et al., 1977; Storlazzi et al., 2003]. This model also revealed that this type of circulation isenhanced by spur-normal waves, increased wave height, weak alongshore currents, increased spur height,and decreased bottom drag.

The classical dynamical basis by which waves drive flow is through changes to the waves from physicalprocesses such as shoaling, refraction, dissipation, etc., which create spatial gradients in radiation stressesand impart a force in the momentum equation [Longuet-Higgins and Stewart, 1964]. Specifically for SAG

Key Points:� Waves over spur and groove

formations induce Lagrangiancirculation cells� Horizontal flow is offshore over spurs

and weak onshore over grooves� Downward flow over spur is likely

due to variation in alongshorebottom stress

Correspondence to:J. S. Rogers,[email protected]

Citation:Rogers, J. S., S. G. Monismith,R. B. Dunbar, and D. Koweek (2015),Field observations of wave-drivencirculation over spur and grooveformations on a coral reef, J. Geophys.Res. Oceans, 120, doi:10.1002/2014JC010464.

Received 19 SEP 2014

Accepted 8 DEC 2014

Accepted article online 16 DEC 2014

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 1

Journal of Geophysical Research: Oceans

PUBLICATIONS

formations, the mechanism for circulation is an imbalance between the cross-shore radiation stress gradientand cross-shore pressure gradient terms in the depth-averaged momentum equations [Rogers et al., 2013].The alongshore variation in radiation stress gradient is primarily due to the local cross-shore slope, whereasthe alongshore variation in mean pressure gradient is primarily due to the local depth; the residual forcingfrom this imbalance accelerates the flow until the bottom stress or nonlinear convection is large enough tobalance it, resulting in alongshore variable flow and the counter-rotating circulation cells [Rogers et al., 2013].

The radiation stress gradient can be recast as a vortex force in the full three-dimensional momentum equa-tions, first proposed by Craik and Leibovich [1976], and refined by others including Kumar et al. [2012]. Thevortex force is the interaction of the Stokes drift with flow vorticity, and is essential in the mechanism forLangmuir circulation. For SAG formations, the three-dimensional velocity structure is unknown but it ishypothesized that due to the coincident Stokes drift and horizontal vorticity in the mean flow, the vortexforce may be important in driving secondary flow.

Another important mechanism capable of creating secondary flow is from lateral (normal to the main flowdirection) periodic variations of bottom stress first proposed by Townsend [1976]. The mechanism of insta-bility is the induction by the normal Reynolds stresses of a pattern of secondary flow, directed from regionsof large stress to ones of small stress; this also induces, by continuity, downward flow over the regions ofhigh stress and upward flow over those of small stress [Townsend, 1976]. It is hypothesized this may be animportant mechanism influencing the secondary flow circulation on SAG formations due to the periodiclarge alongshore differences in bottom roughness between the spur and groove.

Roberts et al. [1977] present, to our knowledge, the only known field measurements of currents on SAGbathymetry based on a single dye release in strong alongshore flow conditions at Grand Cayman, (CaymanIslands). They measured 31 cm/s onshore near-bed velocity in the groove which carried the dye plumeonshore and up and over the spur before being advected alongshore. Beyond the limited data in Robertset al. [1977] which did not resolve the three-dimensional velocity structure, the wave-induced circulationcells predicted by Rogers et al. [2013] have never been observed in the field.

Here we present field observations of wave-induced circulation cells over SAG formations, including theirvertical structure, and discuss several mechanisms consistent with the observed circulation. Based on thenear-bed observations, we discuss why coral growth and development may be enhanced on the spurs. Toaddress these questions, two separate field studies were conducted on Palmyra Atoll in the Central Pacific(section 2.1) and data were processed per accepted methods (section 2.2). Results show wave-induced cir-culation and their vertical velocity profiles (section 3.1), momentum balances (section 3.2), bottom rough-ness characteristics (section 3.3), and near-bed velocity and bottom stress (section 3.4). A discussion onwaves and circulation (section 4.1), mechanism for circulation (section 4.2), implications for coral health (sec-tion 4.3), and conclusions (section 5) follow.

2. Methods

2.1. Field ExperimentPalmyra Atoll (5�520N, 162�050W) is part of the Northern Line Islands of the central equatorial Pacific (Figure1a) and largely because of the absence of acute anthropogenic stress on the ecosystems, its reefs containabundant calcifiers, namely hard corals and crustose coralline algae [Williams et al., 2013] and high growthrates [Koweek et al., 2014]. Two separate field experiments were conducted on the atoll to characterize SAGcirculation cells (Figure 1b). The first experiment, hereafter referred to as SFR12, was conducted from 16 to26 September 2012, on the south forereef in approximately 8–10 m depth (Figures 1b and 1d). The SAG for-mations at SFR12 had approximately 1.8 m high spurs, 15 m average wavelength, 2 m wide grooves, andthe spurs were oriented with an approximately 175� heading offshore, approximately parallel to the largerscale reef slope, which was approximately 15% (Figure 1d). The second experiment, hereafter referred to asNFR13, was conducted from 4 to 9 September 2013, on the northwest corner of the atoll in approximately8–11 m depth (Figures 1b and 1c). The SAG formations at NRF13 had approximately 1.9 m high spurs, 14 maverage wavelength, 1.4 m wide grooves, and the spurs were oriented with an approximately 0� headingoffshore, approximately parallel to the larger scale reef slope, which was approximately 7% (Figure 1c). Livecoral and coralline algae cover was high on the spurs (Figures 2a and 2d), whereas the bottom of thegrooves were typically covered with reef debris, sediment, and fewer live coral colonies (Figures 2b and 2e).

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 2

The NFR13 instrument array consisted of two cross-shore transects of instrumentation covering two spursand two grooves (Figure 1c). Transect A consisted of four sites (A1–A4) measuring velocity and pressure,and transect B consisted of four sites (B1–B4) measuring pressure located approximately 15 m offshorefrom Transect A (Table 1). A deep forereef mooring, Site C1, measuring velocity and pressure was locatedapproximately 115 m downslope from transect A in approximately 20 m water depth. The SFR12 instrumentarray consisted of one cross-shore transect (Transect M) covering one spur and one groove (Figure 1d), con-sisting of two sites (M1 and M2) measuring velocity and pressure (Table 1). A weather station was locatedon the atoll which measured wind speed and direction (RM Young 3002 sensor), and other meteorologicalvariables (Figure 1b).

2.2. Data AnalysisInstantaneous measured velocity data u(u,v,w) were rotated to local bathymetry coordinates of cross-shore(x) and alongshore (y) directions based on the orientation of spurs. The cross-shore coordinate at NFR13 cor-responded to a 0� compass heading, directed offshore (Figure 1c), and cross-shore coordinate at SFR12 cor-responded to a 175� compass heading, directed offshore (Figure 1d). The vertical (z) coordinate is taken asupward from mean sea level (MSL). Time averaging (—) was computed over 15 min intervals for meanvelocity �u, average free surface deviation from MSL, f, and wave statistics. Only velocity data from theADCP/APDs between selected depth ranges were used for analysis (Table 1). The ADV data were combinedwith the ADCP data at sites A2 and A3 to create velocity profiles.

Figure 1. Palmyra Atoll with field experiment location and layout. (a) Palmyra Atoll location in the Central Pacific, (b) overview of the west-ern side of the atoll including the NFR13 experiment location on the north forereef and SFR12 experiment location on the south forereef,and (c) zoom in of NFR13 field experiment showing regional bathymetry and instrument locations with velocity and pressure sensors A1–A4 and C1 (red), and only pressure B1–B4 (blue), and local coordinate system cross-shore (x) and alongshore (y) in relation to true north,and (d) zoom in of SFR12 field experiment showing regional bathymetry and instrument layout with velocity and pressure sensors loca-tions M1 and M2, and local coordinate system cross-shore (x) and alongshore (y) in relation to true south. Note, bathymetry shown in Fig-ures 1c and 1d is regional scale without SAG formations. Image and bathymetry courtesy of NOAA.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 3

Table 1. Experiment Instrumentation for NFR13 and SFR12 Experiments, Sites, Depth, Instrumentation and Sampling Rates

Site Depth h (m) Instrumentation, Sample Rate

NFR13 A1 (Groove) 10.31 1 MHz Nortek Aquadopp 0.5 m bins, 4.0–8.0 MAB, 1 HzRBR Virtuoso, 1 Hz

A2 (Spur) 8.59 600 kHz RDI ADCP 0.4 m bins, 5.3–7.3 MAB, 2 HzNortek ADV, 0.7 MAB, 4 Hz

A3 (Groove) 10.88 1200 kHz RDI ADCP 0.4 m bins, 2.3–9.7 MAB, 2 HzNortek ADV, 0.7 MAB, 4 HzRBR Virtuoso, 1 Hz

A4 (Spur) 9.10 2 MHz Nortek Aquadopp 0.5 m bins, 0.75–7.5 MAB, 1 HzRBR Virtuoso, 1 Hz

B1 (Groove) 10.59 RBR 1050, 1 HzB2 (Spur) 10.22 RBR 1050, 1 HzB3 (Groove) 10.69 RBR 1050, 1 HzB4 (Spur) 10.33 RBR 1050, 1 HzC1 (Deep Forereef) 19.02 1200 kHz RDI ADCP 0.75 m bins, 3.5–17.0 MAB, 2 Hz

RBR 1050, 1 HzSFR12 M1 (Spur) 8.72 1200 kHz RDI ADCP, 0.5 m bins, 2.1–7.5 MAB, 0.67 Hz

RBR 1050, 1 HzM2 (Groove) 10.53 1200 kHz RDI ADCP, 0.5 m bins, 2.1–9.1 MAB, 0.67 Hz

RBR 1050, 1 Hz

Figure 2. Field experiment images and spur and groove bathymetry, for NFR13 experiment (a–c), and SFR12 experiment (d–f). (a) Typicalspur and Station A2 (center) with divers for scale looking onshore, (b) typical groove and Station A3 (center) showing reef debris lookingonshore, and (c) NFR13 spur and groove bathymetry in alongshore direction (Stations A1–A4) showing instrument placement (redsquares), (d) Station M1 showing spur looking alongshore from M2, (e) Station M2 showing groove and reef debris looking alongshorefrom M1, and (f) SFR12 spur and groove bathymetry in alongshore direction (Stations M1–M2) showing instrument placement (redsquares).

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 4

Linear wave theory has been shown to be accurate on rough reefs with steep and complex geometry, andlarge wave amplitudes nearly equal to the mean depth [Monismith et al., 2013]. The use of spectral wavedefinitions (i.e., the superposition many sine waves over many frequencies) is common practice, which doesnot assume a constant wave form and there is some flexibility to include nonlinear wave forms in the for-mulation [Dean and Dalrymple, 1991; Sheremet et al., 2011]. Wave analysis was conducted on pressure p andvelocity data by dividing each 15 min segment into 33 sections of equal length, each with 75% overlap,applying a Hanning window to the segments and computing spectra S fð Þ of frequency f. The rms waveheight Hrms was calculated by Hrms5

Ð 1=51=258Sffdf

h i1=2, where Sff fð Þ is the power spectral density of the free

surface f, calculated from, Sff5Spp cosh kh=ðqgcosh khgÞ� �2

, and Sppðf Þ is the power spectral density of p, qis density, g is gravitational constant, h is depth of the bottom below MSL, hg is the height of the pressuregauge above the bottom, and wave number k is related by the dispersion relation r25gk tanh kh, andradian frequency r52pf 52p=T [Dean and Dalrymple, 1991]. Mean wave period, Tm was calculated based onthe first spectral moment of Sff fð Þ. Mean wave direction hm was computed from the first spectral momentof h fð Þ calculated by, tan 2h fð Þ5Suv fð Þ= Suu fð Þ2Svv fð Þ½ �, where Suu and Svv are the autospectra and Suv is thecospectra of u and v from the ADVs and near-bed ADCP/ADPs bins [Herbers et al., 1999].

The mean Lagrangian velocity uL was calculated by [Andrews and McIntyre, 1978], uL 5uE 1uS , where uE 5

u is the mean measured Eulerian velocity and Stokes drift,

uS 5rH2

rmscosh 2k h1zð Þ8sinh 2kh

k; (1)

was computed spectrally and integrated from 1/5 to 1/25 Hz, and k is the magnitude of wave number vec-tor k [Dean and Dalrymple, 1991]. The Lagrangian depth-averaged mean velocity ULðUL; VL;WLÞ was calcu-lated by combining available data at a given location (ADV/ADCP/ADP), assuming uE 50 at the bottom,linearly interpolating in z and taking the average. To quantify the strength of circulation, we define a cross-shore circulation velocity Uc as [Rogers et al., 2013],

Uc5 UL2hULið Þcos u; (2)

where hULi denotes a spatial average in the alongshore direction to remove the average cross-shore reefflow, and u is the angle between x and y components of UL. In the presence of strong alongshore current(u � p=2) Uc will approach zero; while in the presence of strong cross-shore current (u � 0) Uc willapproach UL2hULi.

Error in the velocity measurements was taken as the measured error from the redundant beam of theADCPs, 1% of the measurement 60.5 cm/s for the ADPs, and 0.5% of the measurement 60.1 cm/s for theADVs. The error was propagated through the calculations per standard error analysis methods [e.g., Emeryand Thomson, 2004].

The approximate depth integrated momentum equations for horizontal flow are given by [e.g., Mei et al., 2005],

@UL

@t1UL � rUL52grf2

1

q f1h� � r � S1sb 2ss½ �; (3)

where S is the radiation stress tensor, sb is the mean bottom stress, and ss is the mean surface stress. Theterms in equation (3) will be referred to from left to right as unsteady (US), convective nonlinear (NL), meanpressure gradient (PG), radiation stress gradient (RS), bottom stress (BT), and surface stress (ST). Equation (3)was evaluated for the NFR13 experiment in the alongshore direction at the midpoint between stations A1–A2 and A2–A3, but was not evaluated between A3 and A4 due to variable local bathymetry and larger dis-tance between sites. Time derivatives were taken using the leapfrog method, while alongshore spatial deriv-atives used a central difference scheme, and cross-shore spatial derivatives used a forward Euler scheme. Aleapfrog scheme was used to compute the alongshore PG due to lack of highly accurate p (RBR) at A2.

For linear waves Syy5Sff n sin 2h11ð Þ2 12

� �, and Sxy5Syx5 1

2 Sffnsin 2h, were computed spectrally and inte-grated over 1/5 to 1/25 Hz, where n5Cg=C, with group velocity Cg and phase speed C [Longuet-Higgins andStewart, 1964]. To evaluate @Sxy=@x, hm at the B stations was assumed equal to the corresponding A station.

The measured velocity can be expressed as u5�u1u01~u, where u0 is the turbulent velocity and ~u is thewave-induced velocity. Assuming motions that are correlated with the free surface are due to waves, and

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 5

those that do not are due to turbulence, S~u ~w 5SufS�wf=Sff, and Su0w05Suw2S~u ~w , and u0w05Ð

Su0w0 df [Benilovand Filyushkin, 1970; Benilov et al., 1973]. The process is similar for the other components of the Reynoldsstress tensor (i.e., u0u0 ) and wave motions (i.e., ~u~u ). The total kinetic energy due to turbulence (TKE) isTKE5 1

2 u0 2 1v 0 2 1w0 2� �

.

Mean bottom stress, sb , was computed from the turbulent Reynolds stress, which are assumed constantwithin the inertial sublayer, [e.g., Reidenbach et al., 2006],

sb 52qu0w 0 (4)

using the measured turbulent velocities (u0) from the ADVs. sb used in equation (3) at the midpointbetween A1 and A2 was assumed equal to the measured sb at A2, and at the midpoint A2 and A3 was takenas the average of sb at A2 and A3. The surface stress was approximated by a typical quadratic drag law,ss 5qaCDau10ju10j, where air density qa51kg m23, wind drag coefficient CDa50:0015, and wind velocity u10

[Smith, 1988].

A common bottom stress parameterization is given by [e.g., Grant and Madsen, 1979; Feddersen et al., 2000],

sb 5qCDujuj; (5)

where u is evaluated near the bed but above the bottom boundary layer, and CD is a nondimensional dragcoefficient which may depend on the flow environment, height above the bed and bottom roughness.Combining equations (4) and (5) gives,

CD52u0w 0

ujuj; (6)

where in environments with low wave and turbulence energy, the denominator is often simplified to �uj�uj,and �u is either the depth-averaged or near-bed velocity, see Rosman and Hench [2011] for a completediscussion.

3. Results

During the NFR13 experiment, the tidal elevation f varied by 0.8 m (Figure 3a). Wave energy was characterizedby two main events, the first on 4–5 September with wide frequency spread, and average Hrms of 0.6 m at 9 speriod, and the second on 8 September with narrower frequency spread and average Hrms of 0.5 m at 11 speriod (Figures 3b–3d). Hrms was slightly higher at deep forereef station C1, but showed only very slight differ-ence between all stations at A and B (Figure 3c). Tm increased with propagation onshore (from C1 to A), butshowed very little alongshore variation (Figure 3d). Shorter period waves were directed on average at 2h, whilethe longer period swell were directed at 1h (Figure 3e), and average wave direction was nearly incident for the4–5 September wave event, while angled at about 20� on 8 September. hm was directed more normally incidentwith onshore propagation consistent with Snell’s law, and hm varied by approximately 10� within sites A1–A4(Figure 3f), essentially the accuracy of the instrument reference frame. The wind speed was directed generally inthe positive x and y directions (approximately to the NW) and its magnitude varied from 0 to 8 m/s (Figure 3g).

During the SFR12 experiment f varied by 0.8 m, (Figure 4a). The waves were characterized by Hrms between0.6 and 1.3 m (Figure 4b), Tm of 8–13 s (Figure 4c), and hm of about zero (incident to cross-shore coordinatex) (Figure 4d). Site M1 (spur) had slightly higher Hrms, and lower Tm compared to Site M2 (groove), but hm

was nearly the same.

3.1. Circulation and Vertical StructureFor the NFR13 experiment, the cross-shore depth-averaged Lagrangian velocity UL varied from 212 to12 cm/s and was generally directed offshore (1) with velocities on the deep forereef (C1) showing similartrends (Figure 5a). UL was generally more positive at A2 (spur) than at A3 (groove) especially on 4, 5, and 7September. The deep forereef (C1) VL varied from 250 to 30 cm/s and was generally positive (to the west)(Figure 5b). At the SAG site A1–A4, VL was smaller in magnitude but followed a similar trend. Two largerscale anomalous flow events occurred on 5 and 8 September with strong flow at C1 directed onshore andnegative alongshore (southeast). WL varied from 24 to 2 cm/s and was generally negative (down) at A2 and

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 6

A4 (spurs) and A1 (groove), and positive at A3 (groove) (Figure 5c). Uc varied from 26 to 6 cm/s and wasnearly always positive at A2 and A4 (spurs) and negative at A1 and A3 (grooves), and was strongest on 4and 5 September (Figure 5d). The horizontal Stokes drift was O(0.1 cm/s) over the experiment duration, andwas thus a small O(10%) fraction of the Lagrangian velocity.

During the SFR12 experiment, UL varied between 26 and 12 cm/s but was generally directed onshore (Fig-ure 4e), VL varied between 226 and 28 cm/s (Figure 4f), and the circulation Uc varied from 21.2 to 1.2 cm/s,and was generally directed offshore at M1 (spur) and onshore at M2 (groove) (Figure 4g). The horizontalStokes drift was O(1 cm/s) over the experiment duration, and was thus a significant fraction of the cross-shore Lagrangian velocity.

A snapshot of the Lagrangian velocity field interpolated in the alongshore direction (NFR13 A stations) at atime of weak alongshore flow and strong Uc, shows characteristic offshore flow cells centered over the spursand weak onshore flow over the groove and at depth (Figure 6). The alongshore and vertical velocitiesshow a strong counterclockwise rotation on the side of the spur (y 5 12 m, h 5 5 m), and there is a sugges-tion of clockwise rotation on the opposite side of the spur (y 5 2 m, h 5 6 m; and y 5 24 m, h 5 5 m).

The characteristic circulation cells shown Figure 6 were present under different flow conditions withstrong Uc, including weak VL and weak UL (Figure 7a), strong VL and weak UL (Figure 7b), weak VL andstrong offshore UL (Figure 7c), strong VL and strong offshore UL (Figure 7d). However, for periods of weakUc, these characteristic circulation cells are not present for different flow conditions including weak VL

and weak UL (Figure 7e), strong VL and weak UL (Figure 7f), weak VL and strong UL (Figure 7g), strong VL

and strong onshore UL (Figure 7h). Where weak and strong refer to below and above the mean, and VL

and UL are taken as the average of stations A1–A4, and Uc is taken as the average of the magnitude fromstations A2 and A3.

In the cross-shore direction, the profile of the rms of the Eulerian velocity uEð Þrms is a complex shape, withNFR13 of larger magnitude than during the SFR12 experiment (Figure 8a). The profile of Stokes velocity

uSð Þrms (from equation (1)) is similar for all NFR13 profiles and very small compared to the Eulerian velocities,while during the SFR12 experiment, Stokes drift is higher over the spur than the groove and of a similarmagnitude to the Eulerian velocity (Figure 8b). The Lagrangian velocity uLð Þrms profile shape is different

Figure 3. Physical forcing of tide, waves, and wind during NFR13 experiment duration, (a) mean free surface f , (b) power spectral densityof surface Sffðf Þ as a function of frequency f at A2, (c) rms wave height Hrms, (d) mean wave period Tm, (e) wave angle h(f) as a function offrequency at A2 (�), (f) mean wave angle hm, and (g) wind velocity U10x (-) and U10y (--). For (a, c, d, and f) blue is A1–A4, red is B1–B4 [notshown for (f)] and black is C1.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 7

between periods of weak and strong Uc for all sites (Figure 8c). For periods of strong Uc, during the NFR13experiment uL rms was much stronger on the spurs about 3 m below the surface (A2, A4), while during theSFR12 experiment the trend is similar but weaker.

Figure 4. Physical forcing of tide, waves, depth-averaged mean Lagrangian velocity UL results and circulation velocity Uc during SFR12experiment duration, (a) mean free surface f , (b) rms wave height Hrms, (c) mean wave period Tm, (d) mean wave direction hm, (e) cross-shore UL, (f) alongshore VL, and (g) circulation Uc. Blue is M1 (spur) and red is M2 (groove). Solid lines (-) are mean values, dotted lines (..)are mean 6 the 95% confidence limit are only shown for (e–g).

Figure 5. Depth-averaged mean Lagrangian velocity UL results and circulation velocity Uc over NFR13 experiment duration, (a) UL, (b) VL,(c) WL, and (d) Uc, Colors are green, blue, red, black, and cyan for A1 (groove), A2 (spur), A3 (groove), A4 (spur), C1 (deep forereef) [notshown for (c) and (d)], respectively. Solid lines (-) are mean values, dotted lines (..) are mean 6 the 95% confidence limit.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8

In the alongshore direction, the profile of vEð Þrms for the bottom half of the water column is similar to a log-layer shape on the spurs, while on the grooves the shape is also like a log-layer but with an offset of aboutthe spur height (Figure 8d). The profiles are similar higher in the water column. The profile of vSð Þrms (fromequation (1)) is similar for all profiles and very small compared to the Eulerian velocities (Figure 8e). Forboth experiments, the profile of vL rms had a similar log-layer-like shape for both strong and weak Uc condi-tions, but strong Uc conditions had a larger magnitude (Figure 8f).

3.2. Momentum BalanceThe significant terms in the depth-averaged cross-shore momentum balance (equation (3)), were of O(1024

m/s2) between nonlinear convective, mean pressure gradient, and radiation stress gradient (RSxy) terms(Table 2) i.e.,

VL@UL

@y� 2g

@f@x

21

q f1h� � @Sxy

@y; (7)

while the unsteady, second nonlinear convective (UL@UL=@x), radiation stress gradient (RSxx), and bottomstress terms were of secondary importance O(1025 m/s2). The significant terms in the depth-averagedalongshore momentum balance (equation (3)), were of O(1024 m/s2) between nonlinear convective, meanpressure gradient, and radiation stress gradient (RSyy) terms (Table 2) i.e.,

VL@VL

@y� 2g

@f@y

21

q f1h� � @Syy

@y; (8)

while the unsteady second nonlinear convective (UL@VL=@x), bottom stress, and radiation stress gradi-ent (RSyx) terms were of secondary importance O(1025 m/s2). Surface stress from wind forcing wasinsignificant at both sites O(1026 m/s2) (Table 2). Because the relative difference in elevation betweenpressure gauges is not known, there may be net bias in the calculated pressure gradient, which is con-stant in time. It is assumed to first order this bias is zero. This assumption is reasonable to first orderbecause the change in wave direction (Figure 3), and the periodic reversal of UL and VL (Figure 5) sug-gests the temporal mean of the mean pressure gradient over the entire measurement period should benear zero. Second, the mean pressure gradient arises as a response to the other forcing terms and isexpected to be of the same order as the other major terms [Rogers et al., 2013], which is the result inTable 2.

Figure 6. Mean Lagrangian velocity uL , in the alongshore (y) and vertical (z) direction showing characteristic spur and groove circulationcells during NFR13 experiment. Arrows indicate alongshore (vL ;wL ) velocity, color shading is cross-shore uL , 1 is offshore and 2 is onshoreflow. Dashed black line is free surface f , solid black is bottom bathymetry, and black x are Stations A1–A4.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 9

3.3. Bottom RoughnessThe bottom coefficient CD was computed at A2 and A3 using the Reynolds stress using equation (6) fromthe ADVs (Table 3). The CD was similar O(0.01) at both sites in the cross-shore direction and in the along-shore direction over the spur (A3), while it was much larger O(0.1) in the alongshore direction over thegroove (A2). Fits to logarithmic profiles were performed for the alongshore Eulerian velocity profiles, [e.g.,Reidenbach et al., 2006] but the results for bottom roughness scale z0 (A3: 0.009 6 0.023 m, A4:0.008 6 0.016 m, C1: 0.02 6 0.04 m) and offset height d (A3: 0.62 6 0.12 m, A4: 0.51 6 0.33 m, C1:0.57 6 1.36) had very large scatter. Use of Grant and Madsen [1979] to remove the effect of waves and pre-dict a physical roughness scale kN (A3: 0.06 6 0.03 m, A4: 0.05 6 0.04 m, C1: 0.12 6 0.12 m) also had veryhigh scatter.

3.4. Near-Bed ResultsDuring the NFR13 experiment, the near-bed average cross-shore Lagrangian velocity uL measured by theADVs was slightly higher in magnitude over the spur (A2) compared to the groove (A3) (Figure 9a), while vL

was significantly higher in magnitude over the spur (Figure 9b). w was similar in magnitude at the two sitesbut was variable over the groove but generally negative (down) over the spur (Figure 9c). The rms of thenear-bed cross-shore wave velocity ~uð Þrms was similar over the spur and the groove for periods of stronger,

Figure 7. Mean Lagrangian velocity uL in alongshore (y) and vertical (z) direction under different flow conditions during NFR13 experiment. For periods of strong Uc with characteristiccirculation cells, (a) weak VL and weak UL, (b) strong VL and weak UL, (c) weak VL and strong UL, (d) strong VL and strong UL; and for periods of weak Uc with little to no circulation cells,(e) weak VL and weak UL, (f) strong VL and weak UL, (g) weak VL and strong UL, (h) strong VL and strong UL. Color scheme and labels are the same as Figure 6, color shading is cross-shoreuL (cm/s).

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 10

directly incident wave forcing (5 September), but larger in the groove for weaker wave forcing (Figure 9d).The measured wave velocity was smaller at both locations than what would be predicted by linear wavetheory (based on the mean Hrms, Tm, and hm). This different was most pronounced over the spur. The u0w 0 ,TKE, and jsb j computed from the wave-separated Reynolds stress, were significantly higher over the spurthan over the groove by up to a factor of four (Figures 9e–9g).

4. Discussion

4.1. Waves and CirculationThe effect of SAG’s on waves appears to be minimal. The NFR13 results show only very small alongshore dif-ferences in wave properties Hrms, Tm, and hm (Figures 3c, 3d, and 3f). For higher wave forcing during theSFR12 experiment, Hrms was slightly higher over the spur (Figure 4b). At both field sites, the alongshore SAGwavelength O(10 m) is much smaller than the surface wavelength O(100 m). As shown by Rogers et al.[2013], for small SAG wavelengths, the effects of diffraction are likely strong, and alongshore differences inwave parameters are therefore minimal. Linear wave theory predicts larger horizontal wave motions overthe spur, but the measured wave velocities were very similar for strong normally incident waves, andsmaller over the spur for weaker incident waves (Figure 9d). Thus, the SAG formations appear to be slowingthe horizontal wave motion over the spur more than what would be predicted by linear wave theory.

For both experiments, the results show the presence of persistent O(1 cm/s) cross-shore circulation velocityUc, directed offshore over the spurs and onshore over the grooves (Figures 4g and 5d) as predicted by themodeling results of Rogers et al [2013]. The period of strongest measured Uc is on 4 September (NFR13),which for that experiment is also the period of highest Hrms, nearly incident waves, and relatively smallalongshore flow VL. Periods of weaker Hrms but directly incident waves (6 September) or high Hrms butangled waves (8 September) show weaker Uc. This is consistent with the modeling results of Rogers et al.[2013], which predicted stronger Uc with high Hrms, directly incident waves (hm 5 0), and weak alongshoreflow (VL). The SFR12 experiment had smaller Uc than the NFR13 experiment likely because while Hrms washigher, the SAG formations were smaller, less well defined (Figure 2), and the alongshore VL was stronger.

Figure 8. Average profiles over depth for rms of mean Eulerian uEð Þrms , Stokes drift uSð Þrms , and Lagrangian uLð Þrms velocity during NFR13and SFR12 experiments. Cross-shore direction (a) uEð Þrms , (b) uSð Þrms , (c) uLð Þrms , and alongshore direction (d) vEð Þrms , (e) vSð Þrms , (f) vLð Þrms .For (c) and (f), periods of above average circulation (jUc j > jUc j) are (2x), and periods of below average circulation (jUc j < jUc j) are (--o).

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 11

For the events on 4, 5, 6, and 7 September, (NFR13) there appears to be a similar growth of Uc at approxi-mately 1 cm/s/h (Figure 5d). Similar growth rate is observed during SFR12 (Figure 4g). Thus, there appearsto be a spin up time for the observed Uc on the order of 1 h for which favorable forcing conditions must bemaintained to accelerate and develop the SAG circulation cells.

The vertical structure of the circulation cells showed strong horizontal offshore flow over the spurs near thesurface, while the grooves had weak horizontal onshore flow (Figure 6). For periods of strong Uc, these cellswere persistent features during periods of different cross and alongshore flow conditions, and were absentduring periods of weak Uc (Figure 7).

The cross-shore Lagrangian flow shows a significant difference in shape in the mean rms profile betweenperiods of strong and weak Uc, i.e., the circulation cells appear to be modifying the velocity profile shape inthe cross-shore direction (Figure 8c). This effect was most pronounced during the NFR13 experiment butwas also observed in the SFR12 experiment. This is in contrast to the alongshore direction which shows alog-layer like flow profile for both strong and weak Uc, but simply of different magnitudes (Figure 8f). Theprofile shape is similar to a log-layer shape on the spurs, while on the grooves the shape is also like a log-layer but with an offset of about the spur height (Figure 8d). We emphasize that because both field sitelocations were near the deeper end of the spurs, the direction of cross-shore flow was offshore over thespur and onshore over the groove. However, as predicted by Rogers et al. [2013] this is likely reversed atshallower depths, but was not investigated in this study.

4.2. Mechanism for CirculationThe cross-shore momentum balance indicates the primary terms were nonlinear convective, mean pressuregradient (RSxy), and radiation stress gradient of O(1024 m/s2). This was the same primary cross-shoremomentum balance obtained by Rogers et al. [2013] for the same relative position on the SAG formations(near the deeper offshore slope). The radiation stress gradient term is the direct result of the shoalingwaves, are most affected by the local bathymetric slope [Rogers et al., 2013], which varies in the alongshoredirection. The mean pressure gradient arises as a response to wave forcing. The residual forcing from theimbalance between the pressure gradient and radiation stress gradient accelerates the flow until the nonlin-ear convection is large enough to balance it, resulting in alongshore variable UL flow and the counter-rotating circulation cells (Figures 5a and 6) consistent with modeling results of Rogers et al. [2013].

The alongshore momentum balance indicates the primary terms were nonlinear convective, mean pressuregradient and radiation stress gradient (RSyy) of O(1023 m/s2). Changes to the waves propagating in the

alongshore direction from refraction, bathymet-ric changes or bottom dissipation create a radia-tion stress gradient (RSyy). As in the cross-shoredirection, the pressure gradient arises inresponse to the forcing and the imbalancebetween the mean pressure gradient and radia-tion stress gradient accelerates the flow until thenonlinear convection is enough to balance it,driving an alongshore VL flow (Figures 5b and 7).

Table 2. Order of Terms in Depth-Averaged Momentum Equations (Equation (3)) From NFR13 Experiment in the Cross-Shore (x) andAlongshore (y) Directionsa

Description

Cross-Shore (x) Alongshore (y)

Term O(m/s2) Term O(m/s2)

Unsteady @UL=@t 1 3 1025 @VL=@t 2 3 1025

Nonlinear convective 1 UL@UL=@x 1 3 1025 VL@VL=@y 2 3 1024

Nonlinear convective 2 VL@UL=@y 2 3 1024 UL@VL=@x 2 3 1025

Mean pressure gradient g@f=@x 2 3 1024 g@f=@y 6 3 1024

Radiation stress gradient 1 b@Sxx=@x 4 3 1025 b@Syy=@y 6 3 1024

Radiation stress gradient 2 b@Sxy=@y 3 3 1024 b@Syx=@x 6 3 1026

Bottom stress bsbx 2 3 1025 bsby 1 3 1025

Surface stress bssx 2 3 1026 bssy 2 3 1026

aNote: b51= q f1h� �� �

. Order O is the average of the absolute value of the term over the experiment duration.

Table 3. Bottom Drag Coefficient CD Results From NFR13 Experi-ment From Near-Bed ADV Measurements in Cross-Shore (x) andAlongshore (y) Directionsa

A2 (Spur) A3 (Groove)

CD (x) 0.015 6 0.001 0.011 6 0.001CD (y) 0.0072 6 0.0007 0.10 6 0.007

aNote: Reference height zref 5 0.7 m, confidence intervals are 1standard deviation.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 12

For both cross and alongshore directions, the nonlinear convective terms are important at this site due tothe large variability in bathymetry from the SAG formations and rough reef. The largest errors in themomentum balance are likely from the pressure gradient terms due to instrument accuracy in p measure-ment and the unknown net constant bias between gauges (assumed to be zero). The second largest errorsin the momentum balance are likely the smaller nonlinear convective terms (cross-shore UL@UL=@x, along-shore UL@VL=@x), which were likely larger than the calculated value due to the approximation of @ =@xbetween Transect A and C.

The characteristic shape of the flow field appears analogous to Langmuir circulation cells but of oppositerotation in the yz plane (Figure 6). Since the vertical vorticity xz at h 5 4 m, y 5 11 m is positive, and uS isnegative, the horizontal vortex force, uS3xz [Craik and Leibovich, 1976] is directed toward positive y, con-trary to the observed circulation in the yz plane. For waves and currents opposed, the vortex force is stabiliz-ing [Leibovich, 1983], and thus it appears unlikely the vortex force is the mechanism for the observedrotation in the yz plane. However, only limited studies have been conducted on Langmuir circulation withopposing waves and currents and to our knowledge no studies have been conducted on alongshore peri-odic bathymetry [Thorpe, 2004]. Additionally, further onshore where waves and currents are aligned [Rogerset al., 2013], the vortex force would be destabilizing [Leibovich, 1983]. Thus, further study seems warranted.

The near-bed cross-shore Reynolds stress u0w 0 is nearly always higher on the spur than on the groove (Fig-ure 9e), and the near-bed mean vertical motions WL are nearly always directed down over the spur and upover the groove (Figure 9c), resulting in the larger secondary flow pattern seen in Figure 6, consistent withthe lateral stress mechanism and secondary flow pattern proposed by Townsend [1976]. On 7 September,u0w 0 is nearly the same on the spur and groove corresponding to very little difference in WL between thespur and groove also consistent with this mechanism. Therefore, the lateral stress mechanism of Townsend[1976] seems the most likely mechanism for the observed downward flow over the spur and upward flowover the groove. However, to our knowledge no studies exist to examine this mechanism in a similar flowenvironment and thus further study seems warranted.

The computed bottom drag coefficient values, CD were similar to values computed for other reefs [Rosmanand Hench, 2011]. CD on the spur in both directions and over the groove in the cross-shore direction are of

Figure 9. Near-bed mean Lagrangian velocity and bottom stress results from ADVs (h 5 0.7 m) during NFR13 experiment duration, (a)cross-shore uL (cm/s), (b) alongshore vL (cm/s), (c) vertical wL (cm/s), (d) cross-shore ~uð Þrms (cm/s), (e) cross-shore Reynolds stress u0w0

(m2/s2), (f) turbulent kinetic energy TKE (m2/s2), and (g) bottom shear stress magnitude jsb j (Pa). Colors are blue and red for A2 (spur) andA3 (groove), respectively. For (d), result from linear wave theory is cyan and magenta for A2 (spur) and A3 (groove), respectively.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 13

similar magnitude, while over the groove in the alongshore direction is an order of magnitude larger. Thelarger CD is likely the result of the SAG morphology affecting the flow profile over the groove in the along-shore direction. This effect is seen in the alongshore Eulerian velocity profile which shows low velocity waterover the grooves near the bed, while higher up in the water column the profiles are similar over both thespur and grooves (Figure 8d). These results suggest that the form drag from the SAG morphology in thealongshore direction may be dominant over differences in the frictional drag from benthic cover (i.e., coralversus debris/sand). These results for CD are local measurements, but the net average CD over the largerscale SAG morphology O(100 m) requires further study.

4.3. Implications for Coral HealthThe near-bed velocity, turbulence, and bottom shear stress are of particular interest because they directlyaffect organisms on the bed. Differences in these parameters between the spur and groove may illuminatewhy corals thrive on the spurs and not in the grooves, beyond the influence of light levels that are generallyhigher on the spur. While the cross-shore near-bed velocity was slightly higher on the spur (Figure 9a), thealongshore near-bed velocity was much higher (Figure 9b) during the NFR13 experiment. The vertical velocitywas also directed down over the spur, while directed up over the groove (Figures 6 and 9c). There was alsoincreased turbulence kinetic energy over the spur (Figure 9f). Corals subjected to stronger water motion havegreater mass transfer, including increased nutrient uptake rates [Atkinson and Bilger, 1992; Thomas and Atkin-son, 1997; Atkinson et al., 2001], photosynthetic production and calcification [Dennison and Barnes, 1988], andparticle capture [Genin et al., 2009]. Since corals living on the spurs are the recipients of much higher along-shore flows and variable vertical flow from the surface providing increased ‘‘food’’ supply, and increased turbu-lent motions providing increased contact, they may have a net advantage to corals living in the grooves.

The bottom shear stress was also significantly higher and more variable on the spur than the groove withan average of 0.37 6 0.21 and 0.17 6 0.07 Pa, respectively, during the NFR13 experiment (Figure 9g). Assum-ing incipient motion based on the shields parameter [e.g., Julien, 2010], this would correspond to flow capa-ble of suspending sediments smaller than very fine gravel on the spur and coarse sand on the grooveduring the relatively small wave conditions observed during the NFR13 experiment. During higher waveevents, bottom stress would be much higher. Since the spur has much higher mean flows primarily in thealongshore direction, and bottom slope is toward the grooves, sediment suspended by the bottom stresswould be shed from the spurs toward the grooves where it could accumulate. This sediment and debrisaccumulation was present at the study sites (Figures 2b and 2e), and has been observed as a primary fea-ture of SAG formations around the world [Rogers et al., 2013]. Over time, the sediment in the grooves is car-ried downslope [Storlazzi et al., 2003]. Lower debris and sediment accumulation on the spurs relative to thegrooves would be a significant advantage for recruitment and coral growth [Buddemeier and Hopley, 1988;Acevedo et al., 1989; Rogers, 1990; Fortes 2000; Fabricius, 2005]. During large wave events, the bottom stresswould likely be much greater than what was observed during the NFR13 study period (such as duringSFR12 experiment); potentially subjecting the coral on the spur to proportionally higher bottom shear stressand perhaps breakage. However, the magnitude of this effect remains unclear and requires further study.

5. Conclusions

The results from two separate field studies of SAG formations on Palmyra Atoll show the effect of SAG for-mations on waves was small, and there was a persistent O(1 cm/s) depth-averaged Lagrangian circulation(Uc) of offshore flow over the spurs and onshore flow over the grooves. This circulation was stronger forlarger, directly incident waves and low alongshore flow conditions. There also appeared to be a spin-uptime for the observed Uc on the order of 1 h for which favorable forcing conditions must be maintained toaccelerate and develop the SAG circulation cells. These are the first field observations of SAG hydrodynam-ics and confirm the modeling results from Rogers et al. [2013]. The primary cross and alongshore momen-tum balances were between the pressure gradient, radiation stress gradient and nonlinear convectiveterms. The vertical structure of these circulation cells was previously unknown and the results show a com-plex horizontal offshore Lagrangian flow over the spurs near the surface driven by alongshore variability inradiation stress gradients consistent with Rogers et al. [2013]. Vertical flow was downward over the spur andupward over the groove, likely driven by alongshore differences in bottom stress [Townsend, 1976] and notby vortex forcing [Craik and Leibovich, 1976]. The bottom drag coefficients were similar to values found on

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 14

other reefs; and were enhanced over the groove in the alongshore direction. Beyond the influence of lightlevels that are generally higher on the spur, we suggest that the conditions for coral recruitment andgrowth appear to be more favorable on the spur than the groove due to (1) higher ‘‘food’’ supply fromhigher mean alongshore velocity, downward vertical velocity, and higher turbulence, and (2) lower sedi-ment accumulation due to higher and more variable bottom shear stress.

The present study was conducted near the deeper end of the SAG formations. At shallower depths, thedirection of the depth-averaged cross-shore circulation (Uc) is likely reversed [Rogers et al., 2013], but thiseffect and the three-dimensional Lagrangian velocity structure at these shallower depths remains unknownand requires further study. The observed similarity to Langmuir circulation but with opposite rotation, andthe possible importance of the vortex force and lateral variation of stress mechanism would warrant furtheranalytical or modeling work. Additionally, SAG hydrodynamics under large wave conditions, as well asinvestigation into the sediment transport through the grooves, also warrants further inquiry.

ReferencesAcevedo, R. J., J. Morelock, and R. A. Olivieri (1989), Modification of coral reef zonation by terrigenous sediment stress, Palaios, 4, 92–100,

doi:10.2307/3514736.Andrews, D., and M. McIntyre (1978), An exact theory of nonlinear waves on a Lagrangian-mean flow, J. Fluid Mech., 89, 609–646, doi:

10.1017/S0022112078002773.Atkinson, M. J., and R. W. Bilger (1992), Effect of water velocity on phosphate uptake in coral reef-flat communities, Limnol. Oceanogr., 37,

273–279, doi:10.4319/lo.1992.37.2.0273.Atkinson, M. J., J. Falter, and C. Hearn (2001), Nutrient dynamics in the biosphere 2 coral reef mesocosm: Water velocity controls NH4 and

PO4 uptake, Coral Reefs, 20, 341–346, doi:10.1007/s00338-001-0184-7.Benilov, A. Y., and B. N. Filyushkin (1970), Application of methods of linear filtration to an analysis of fluctuations in the surface layer of the

sea, Izv. Atmos. Oceanic Phys., 6, 810–819.Benilov, A. Y., O. Kouznetsov, and G. Panin (1973), On the analysis of wind wave-induced disturbances in the atmospheric turbulent surface

layer, Boundary Layer Meteorol., 6, 269–285, doi:10.1007/BF00232489.Blanchon, P., and B. Jones (1997), Hurricane control on shelf-edge-reef architecture around Grand Cayman, Sedimentology, 44, 479–506,

doi:10.1046/j.1365-3091.1997.d01-32.x.Buddemeier, R. W., and D. Hopley (1988), Turn-ons and turn-offs: Causes and mechanisms of the initiation and termination of coral reef

growth, in Sixth International Coral Reef Symposium, pp. 253–261, International Society for Reef Studies, Townsville, Australia.Chappell, J. (1980), Coral morphology, diversity and reef growth, Nature, 286, 249–252, doi:10.1038/286249a0.Craik, A., and S. Leibovich (1976), A rational model for Langmuir circulations, J. Fluid Mech., 73, 401–426, doi:10.1017/S0022112076001420.Dean, R. G., and R. A. Dalrymple (1991), Water Wave Mechanics for Engineers and Scientists, Adv. Ser. on Ocean Eng., vol. 2. World Sci., Hack-

ensack, N. J.Dennison, W. C., and D. J. Barnes (1988), Effect of water motion on coral photosynthesis and calcification, J. Exp. Mar. Biol. Ecol., 115, 67–77,

doi:10.1016/0022-0981(88)90190-6.Emery, W. J., and R. E. Thomson (2004), Data Analysis Methods in Physical Oceanography, 2nd ed., 638 pp., Elsevier, Amsterdam.Fabricius, K. E. (2005), Effects of terrestrial runoff on the ecology of corals and coral reefs: Review and synthesis, Mar. Pollut. Bull., 50(2),

125–146, doi:10.1016/j.marpolbul.2004.11.028.Feddersen, F., R. T. Guza, S. Elgar, and T. H. C. Herbers (2000), Velocity moments in alongshore bottom stress parameterizations, J. Geophys.

Res., 105(C4), 8673–8686, doi:10.1029/2000JC900022.Fortes, M. (2000), The effects of siltation on tropical coastal ecosystems, in Oceanographic Processes of Coral Reefs, edited by E. Wolanski,

pp. 93–112, CRC Press, Boca Raton, Fla.Genin, A., S. G. Monismith, M. A. Reidenbach, G. Yahel, and J. R. Koseff (2009), Intense benthic grazing in a coral reef, Limnol. Oceanogr.

Methods, 54(3), 938–951, doi:10.4319/lo.2009.54.3.0938.Grant, W. D., and O. S. Madsen (1979), Combined wave and current interaction with a rough bottom, J. Geophys. Res., 84(C4), 1797–1808,

doi:10.1029/JC084iC04p01797.Herbers, T., S. Elgar, R. Guza (1999), Directional spreading of waves in the nearshore, J. Geophys. Res., 104(C4), 7683–7693, doi:10.1029/

1998JC900092.Julien, P. (2010), Erosion and Sedimentation, 392 pp., Cambridge Univ. Press, N. Y.Koweek, D., R. B. Dunbar, J. S. Rogers, G. J. Williams, N. Price, D. Mucciarone, and L. Teneva (2014), Environmental and ecological controls of

coral community metabolism on Palmyra Atoll, Coral Reefs, 1–13, doi:10.1007/s00338-014-1217-3.Kumar, N., G. Voulgaris, J. Warner, and M. Olabarrieta (2012), Implementation of the vortex force formalism in the coupled ocean-

atmosphere-wave-sediment transport (COAWST) modeling system for inner shelf and surf zone applications, Ocean Modell., 47, 65–95,doi:10.1016/j.ocemod.2012.01.003.

Leibovich, S. (1983), The form and dynamics of Langmuir circulations, Annu. Rev. Fluid Mech., 15, 391–427, doi:10.1146/annurev.fl.15.010183.002135.

Longuet-Higgins, M., and R. Stewart (1964), Radiation stresses in water waves; a physical discussion, with applications, Deep Sea Res. Ocean-ogr. Abstr., 11(4), 529–562, doi:10.1016/0011-7471(64)90001-4.

Masselink, G., and M. G. Hughes (2003), Introduction to Coastal Processes and Geomorphology, 354 pp., Oxford Univ. Press, N. Y.Mei, C., M. Stiassnie, and D. Yue (2005), Theory and Applications of Ocean Surface Waves, Adv. Ser. on Ocean Eng., vol. 23, World Sci., Hacken-

sack, N. J.Monismith, S., L. Herdman, S. Ahmerkamp, and J. Hench (2013), Wave transformation and wave-driven flow across a steep coral reef, J.

Phys. Oceanogr., 43, 1356–1379, doi:10.1175/JPO-D-12-0164.1.Munk, W. H., and M. C. Sargent (1954), Adjustment of Bikini Atoll to ocean waves, U.S. Geol. Surv. Prof. Pap., 260-C, 275–280, U.S. Govern.

Print. Off., Washington, D. C.

AcknowledgmentsData from this study has beendeposited at the NOAA NationalOceanographic Data Center and canbe obtained there (accession0123612). This project was funded bythe Gordon and Betty MooreFoundation, and supported with anNDSEG Fellowship to JSR (U.S.Department of Defense, Office ofNaval Research, 32 CFR 168a), and anNSF Graduate Research Fellowship toDK. We wish to acknowledge able fieldassistance from Mallory Barkdull, HankLynch, David Mucciarone, and LidaTeneva. Helpful discussions with FalkFeddersen and Curt Storlazzi andcomments from two anonymousreviewers improved this manuscript.Logistical and operational support onPalmyra Atoll was provided by theNature Conservancy; this research wasconducted under US Fish and WildlifeService Special Use Permit 12533-13030, and this is Palmyra AtollResearch Consortium contributionPARC-0113.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 15

Reidenbach, M. A., S. G. Monismith, J. R. Koseff, G. Yahel, and A. Genin (2006), Boundary layer turbulence and flow structure over a fringingcoral reef, Limnol. Oceanogr. Methods, 51(5), 1956–1968, doi:10.4319/lo.2006.51.5.1956.

Roberts, H. H. (1974), Variability of reefs with regard to changes in wave power around an island, paper presented at the 2nd InternationalCoral Reef Symposium, International Society for Reef Studies, Brisbane, Australia.

Roberts, H. H., S. P. Murray, and J. N. Suhayda (1977), Evidence for strong currents and turbulence in a deep coral reef groove, Limnol. Oce-anogr., 22(1), 152–156, doi:10.4319/lo.1977.22.1.0152.

Rogers, C. S. (1990), Responses of coral reefs and reef organisms to sedimentation. Mar. Ecol. Prog. Ser., 62, 185–202, doi:10.3354/meps062185.

Rogers, J. S., S. G. Monismith, F. Feddersen, and C. D. Storlazzi (2013), Hydrodynamics of spur and groove formations on a coral reef, J. Geo-phys. Res. Oceans, 118, 3059–3073, doi:10.1002/jgrc.20225.

Rosman, J. H., and J. L. Hench (2011), A framework for understanding drag parameterizations for coral reefs, J. Geophys. Res., 116, C08025,doi:10.1029/2010JC006892.

Sheremet, A., J. Kaihatu, S. Su, E. Smith, and J. Smith (2011), Modeling of nonlinear wave propagation over fringing reefs, Coastal Eng.,58(12), 1125–1137, doi:10.1016/j.coastaleng.2011.06.007.

Smith, S. D. (1988), Coefficients for sea surface wind stress, heat flux, and wind profiles as a function of wind speed and temperature, J.Geophys. Res., 93(C12), 15,467–15,472, doi:10.1029/JC093iC12p15467.

Storlazzi, C., J. Logan, and M. Field (2003), Quantitative morphology of a fringing reef tract from high-resolution laser bathymetry: SouthernMolokai, Hawaii, Geol. Soc. Am. Bull., 115(11), 1344–1355, doi:10.1130/B25200.1.

Storlazzi, C., E. Brown, M. Field, K. Rogers, and P. Jokiel (2005), A model for wave control on coral breakage and species distribution in theHawaiian Islands, Coral Reefs, 24, 43–55, doi:10.1007/s00338-004-0430-x.

Thomas, F., and M. Atkinson (1997), Ammonium uptake by coral reefs: Effects of water velocity and surface roughness on mass transfer,Limnol. Oceanogr., 42(1), 81–88, doi:10.4319/lo.1997.42.1.0081.

Thorpe, S. A. (2004), Langmuir circulation, Annu. Rev. Fluid Mech., 36, 55–79, doi:10.1146/annurev.fluid.36.052203.071431.Townsend, A. A. (1976), The Structure of Turbulent Shear Flow, 2nd ed., 429 pp., Cambridge Univ. Press, N. Y.Williams, G. J., J. E. Smith, E. J. Conklin, J. M. Gove, E. Sala, and S. A. Sandin (2013), Benthic communities at two remote Pacific coral reefs:

Effects of reef habitat, depth, and wave energy gradients on spatial patterns, Peer J, 1, e81, doi:10.7717/peerj.81.

Journal of Geophysical Research: Oceans 10.1002/2014JC010464

ROGERS ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 16