On-shelf transport of slope water lenses within the seasonal pycnocline

Similar documents
Variations of Kuroshio Intrusion and Internal Waves at Southern East China Sea

Processes Coupling the Upper and Deep Ocean on the Continental Slope

Local generation of internal solitary waves in an oceanic pycnocline

Internal Waves in the Vicinity of the Kuroshio Path

Internal Wave Generation and Scattering from Rough Topography

On the interaction between internal tides and wind-induced near-inertial currents at the shelf edge

Upper Ocean Circulation

Generation and Evolution of Internal Waves in Luzon Strait

Mean Stream-Coordinate Structure of the Kuroshio Extension First Meander Trough

Ocean Mixing and Climate Change

A Model Study of Internal Tides in Coastal Frontal Zone*

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts

3. Midlatitude Storm Tracks and the North Atlantic Oscillation

Submesoscale Routes to Lateral Mixing in the Ocean

Modeling the Columbia River Plume on the Oregon Shelf during Summer Upwelling. 2 Model

Features of near-inertial motions observed on the northern South. China Sea shelf during the passage of two typhoons

Baroclinic Rossby waves in the ocean: normal modes, phase speeds and instability

Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability

General Comment on Lab Reports: v. good + corresponds to a lab report that: has structure (Intro., Method, Results, Discussion, an Abstract would be

Variability in the Slope Water and its relation to the Gulf Stream path

that individual/local amplitudes of Ro can reach O(1).

Upper-Ocean Processes and Air-Sea Interaction in the Indonesian Seas

Applying Basin-Scale HyCOM Hindcasts in Providing Open Boundary Conditions for Nested High-Resolution Coastal Circulation Modeling

A process study of tidal mixing over rough topography

MARINE RESEARCH. Journal of. Water-mass transformation in the shelf seas. Volume 68, Number 2

ROSSBY WAVE PROPAGATION

National Oceanography Centre. Research & Consultancy Report No. 36

SAMS Gliders: Research Activities

Hydrodynamics in Shallow Estuaries with Complex Bathymetry and Large Tidal Ranges

Island Wakes in Shallow Water

Erratic internal waves at SIO Pier. data and wavelet analysis courtesy of E. Terrill, SIO

Modeling and Parameterizing Mixed Layer Eddies

Laboratory Modeling of Internal Wave Generation in Straits

Down-welling circulation of the northwest European continental shelf: A driving mechanism for the continental shelf carbon pump

Buoyancy-forced circulations in shallow marginal seas

Goals of this Chapter

Coastal Antarctic polynyas: A coupled process requiring high model resolution in the ocean and atmosphere

Water mass formation, subduction, and the oceanic heat budget

A large-amplitude meander of the shelfbreak front during summer south of New England: Observations from the Shelfbreak PRIMER experiment

SMS 303: Integrative Marine

Alexander Barth, Aida Alvera-Azc. Azcárate, Robert H. Weisberg, University of South Florida. George Halliwell RSMAS, University of Miami

Physical Oceanography of the Northeastern Chukchi Sea: A Preliminary Synthesis

APPENDIX B PHYSICAL BASELINE STUDY: NORTHEAST BAFFIN BAY 1

Comparison Figures from the New 22-Year Daily Eddy Dataset (January April 2015)

Climate impact on interannual variability of Weddell Sea Bottom Water

Journal of Marine Systems

Donald Slinn, Murray D. Levine

Instability of a coastal jet in a two-layer model ; application to the Ushant front

Small scale mixing in coastal areas

A twenty year reversal in water mass trends in the subtropical North Atlantic

Tidally Induced Cross-frontal Mean Circulation: A Numerical Study 1

SIO 210 Problem Set 2 October 17, 2011 Due Oct. 24, 2011

Numerical Experiment on the Fortnight Variation of the Residual Current in the Ariake Sea

Modeling the Lateral Circulation in Straight, Stratified Estuaries*

DIAMIX an experimental study of diapycnal deepwater mixing in the virtually tideless Baltic Sea

The propagating response of coastal circulation due to wind relaxations along the central California coast

Energy Budget of Nonlinear Internal Waves near Dongsha

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming

the only tropical inter-ocean exchange site (~15 Sv) transports heat and freshwater from Pacific into Indian Ocean pressure gradient between Pacific

The Role of Advection in Determining the Temperature Structure of the Irish Sea

Ocean Dynamics. The Great Wave off Kanagawa Hokusai

Mesoscale Processes over the Shelf and Slope in SW06

Spreading of near-inertial energy in a 1/12 model of the North Atlantic Ocean

Energy Dissipation Studies in the South China Sea

Capabilities of Ocean Mixed Layer Models

Contents. Parti Fundamentals. 1. Introduction. 2. The Coriolis Force. Preface Preface of the First Edition

Lecture 9: Tidal Rectification, Stratification and Mixing

1/27/2010. With this method, all filed variables are separated into. from the basic state: Assumptions 1: : the basic state variables must

SIO 210 Introduction to Physical Oceanography Mid-term examination Wednesday, November 2, :00 2:50 PM

The Large Scale Response of the Upper Ocean to Atmospheric Forcing During TOGA-COARE

196 7 atmospheric oscillations:

An Analysis of 500 hpa Height Fields and Zonal Wind: Examination of the Rossby Wave Theory

Tidal and spring-neap variations in horizontal dispersion in a partially mixed estuary

Energy flux of nonlinear internal waves in northern South China Sea

For example, for values of A x = 0 m /s, f 0 s, and L = 0 km, then E h = 0. and the motion may be influenced by horizontal friction if Corioli

Homework 5: Background Ocean Water Properties & Stratification

Impact of frontal eddy dynamics on the Loop Current variability during free and data assimilative HYCOM simulations

REPORT DOCUMENTATION PAGE

psio 210 Introduction to Physical Oceanography Mid-term examination November 3, 2014; 1 hour 20 minutes Answer key

An Introduction to Coupled Models of the Atmosphere Ocean System

A Modeling Study of Eulerian and Lagrangian Aspects of Shelf Circulation off Duck, North Carolina

Estimates of Diapycnal Mixing Using LADCP and CTD data from I8S

The North Atlantic Oscillation: Climatic Significance and Environmental Impact

Sediment Transport at Density Fronts in Shallow Water: a Continuation of N

DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Lateral Mixing

Influence of wind direction, wind waves, and density stratification upon sediment transport in shelf edge regions: The Iberian shelf

Lecture 1. Equations of motion - Newton s second law in three dimensions. Pressure gradient + force force

Meridional coherence of the North Atlantic meridional overturning circulation

Low frequency variability on the continental slope of the southern Weddell Sea

Impact of Alaskan Stream eddies on chlorophyll distribution in the central subarctic North Pacific* Hiromichi Ueno 1,

Hepeng Zhang Ben King Bruce Rodenborn

Quantify Lateral Dispersion and Turbulent Mixing by Spatial Array

Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability

Satellite-derived environmental drivers for top predator hotspots

Monthly climatology of the continental shelf waters of the South Atlantic Bight

Chapter 1. Introduction

Winds and Global Circulation

SIO 210 Introduction to Physical Oceanography Mid-term examination November 5, 2012; 50 minutes Answer key

DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited.

The Taiwan-Tsushima Warm Current System: Its Path and the Transformation of the Water Mass in the East China Sea

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL. 39,, doi:10.1029/2012gl051388, 2012 On-shelf transport of slope water lenses within the seasonal pycnocline J. Hopkins, 1 J. Sharples, 1,2 and J. M. Huthnance 1 Received 19 February 2012; revised 28 March 2012; accepted 28 March 2012; published 26 April 2012. [1] We show that discrete lenses of anomalously highsalinity water, originating from the shelf edge and trapped within the seasonal pycnocline, are advected 100 km or more onto the Celtic Sea continental shelf. We propose that the lenses are created by increased diapycnal mixing at the shelf edge associated with breaking high-frequency internal wave packets. Quasi-synoptic hydrography sections show the lenses to be 3 5 km wide, their temporal persistence confirmed by moored instrumentation and a series of CTD casts. Estimates of the propagation speed of these features (0.020 m s 1 ) compare favorably with the magnitude of observed residual currents. Residual current variability within the pycnocline is dominated by vertical structures most consistent with the second baroclinic mode. The residual flow is therefore thought to be predominantly driven by non-linear second mode internal tidal waves. These are observations of a shelf edge exchange process not previously identified. Citation: Hopkins, J., J. Sharples, and J. M. Huthnance (2012), On-shelf transport of slope water lenses within the seasonal pycnocline, Geophys. Res. Lett., 39,, doi:10.1029/2012gl051388. 1 National Oceanography Centre, Liverpool, UK. 2 Department of Earth and Ocean Sciences, University of Liverpool, Liverpool, UK. Corresponding Author: J. Hopkins, National Oceanography Centre, 6 Brownlow St., Liverpool L3 5DA, UK. (j.hopkins@noc.ac.uk) Published in 2012 by the American Geophysical Union. 1. Introduction [2] The exchange of water between the deep ocean and continental shelf plays an important role in the cycling of carbon and nutrients [Liu et al., 2010], and in flushing of fresh water and sediments from the shelf [McCave et al., 2001]. Here we present observations of a new shelf edge exchange process: lenses of high-salinity slope water, trapped within the pycnocline, being transferred 100 km or more onto the continental shelf. Historical data confirms that salinity anomalies within the pycnocline are a common occurrence, but never before has this signature been reliably captured in high resolution hydrography sections. [3] The observations were made in the Celtic Sea, a 500 km wide section of the NW European shelf (Figure 1a). Depths on the shelf are typically 100 150 m and increase rapidly to more than 2000 m beyond the shelf edge. Internal tides are prominent during the stratified summer months, having peak-to-trough amplitudes of >50 m at the shelf edge during spring tides [Pingree et al., 1986]. With wavelengths of the order 35 km, they may coherently propagate 170 km or more on-shelf [Inall et al., 2011]. [4] Ocean-shelf water exchange in the Celtic Sea is thought to be dominated by drainage in the bottom Ekman layer, nonlinearity of the internal tide and prevailing winds driving a cross-shelf surface Ekman transport [Huthnance et al., 2009]. Here we propose that residual currents within the pycnocline, driven by non-linear mode two internal waves, are responsible for on-shelf advection of discrete lenses of high salinity water generated at the shelf edge by enhanced vertical internal tidal mixing. The salinity anomalies described here have similar characteristics to the salty intrusions observed over the Middle Atlantic Bight (MAB) for which a number of formation mechanisms have been hypothesized, including wind forcing, warm-core rings, and baroclinic pressure gradients [Lentz, 2003]. Owing to the contrasting dynamics of the Celtic Sea and MAB however our proposed formation mechanism is different. 2. Observations [5] Observations were made over the 5 13 June 2010 in the Celtic Sea when thermally driven summer time stratification was well developed, at a location 100 km from the shelf break (Figures 1a and 1b). Despite wind speeds peaking at over 10 m s 1 on three separate occasions the thermocline was not broken down. A Seabird 911 plus CTD provided water column profiles around two mooring sites, IM1 and IM3, where bottom mounted 150 khz ADCPs and thermistor chains were deployed. Scanfish, an undulating vehicle fitted with a SeaBird 911 plus CTD, towed at 7 knots behind the ship, provided sawtooth-like up and down casts between the surface and 90 m. A complete up and down cycle was completed over a distance of 1.3 km. In total, over 1000 km of Scanfish data was collected covering a (140 km) 2 area (Figure 1b). Shear micro-structure measurements were made around the moorings using a Rockland Scientific VMP-500 micro-structure profiler. [6] Tidal analysis shows barotropic currents to be dominated by a clockwise semi-diurnal tide with a semi-major axis amplitude of 0.35 m s 1. The major axis is inclined 45 counter-clockwise from due east, a direction orthogonal to the shelf-edge, and we define across-shelf currents as being along this axis. Residual currents were calculated by subtraction of the fitted barotropic tide (with M 2,S 2,N 2, K 1,O 1 and M 4 constituents). 3. Saline Pycnocline Lenses [7] A 90 km long Scanfish section normal to the shelf edge (Figure 1c) reveals a train of discrete lenses of anomalously high salinity water occupying the entire depth of the pycnocline (25 30 m). The lens width increases from 3 to 5 km with distance on-shelf. The salinity maximum at the centre of each lens is 0.04 0.09 greater than the salinity in surrounding surface and bottom waters. Transects in an along-shelf direction reveal similar isolated 3 5 km wide 1of6

Figure 1. (a) Bathymetry of the NW European shelf (m) with 200 m shelf edge isobath contoured. (b) SST ( C) composite for June 2010 around survey area (box in Figure 1a). Thick white line marks the Scanfish transect shown in Figure 1c. Black dots show the location of lenses over the survey area. IM1 and IM3 moorings are located within the open white circle. (c) Salinity along Scanfish transect. Contours are the top (s q = 26.4) and bottom (s q = 27.2) of the pycnocline. (d) Maximum (black) and mean (gray) salinity within the pycnocline. Vertical arrow indicates the range of peak pycnocline salinities from the CTD casts. Values used to estimate the salt flux are taken from the two dots. anomalies confirming that these are discrete, approximately symmetric features rather than elongated along-shelf streaks. By identifying local maxima in the along-track pycnocline salinity anomaly, we estimate that approximately 160 lenses were observed in the 1000 km of Scanfish data collected (Figure 1b). [8] The vertical structure of the lenses was captured by a series of 26 CTD casts taken over 7.6 days. The salinity maximum within the pycnocline is 35.41 0.01 (mean 1s.d) and lies along the 26.9 0.1 isopycnal (at a depth of 30.7 4.2 m). The lens is 29.3 5.1 m thick and has a maximum salinity of 0.05 and 0.07 greater than surface and bottom waters respectively (Figure 2a). These casts were taken over a range of different tidal states, the first approximately three days before neaps, and the last four days postneap. The temporal persistence of the anomaly indicates that it was a stable feature, at least over the period of observation. The range of maximum pycnocline salinity values captured in the CTD casts (35.40 35.42) matches the peak salinities observed by Scanfish within and between lenses as it passed closest to the mooring site (Figure 1d). This provides evidence that anomalies are propagating past a fixed point. 4. A Pulsed Shelf Edge Salt Source [9] Water with a salinity of 35.4 found mid-shelf is anomalous and must originate from the shelf edge where typical salinities are 35.65. The pulsed nature of the lenses indicates that there is not a continuous flow of saline water into the pycnocline, rather, that there is a regular, pulsed source. [10] The generation of an M 2 internal tidal wave at the shelf edge as the pycnocline is depressed during off-shelf barotropic tidal flow leads to the formation of steep, higher frequency non-linear waves, which form on-shelf propagating packets [Pingree and New, 1995]. The associated strong current shear leads to increased localized turbulence and vertical mixing [Rippeth and Inall, 2002] that is responsible for the cool band of surface water along the shelf edge (Figure 1b). Given that these 1 2 km wavelength waves typically form in packets of 2 8 [New and Pingree, 1990], the local horizontal scale of each mixing event is approximately 2 16 km. We propose therefore that increased diapycnal mixing at the shelf edge, associated with breaking internal wave packets, generated at a semi-diurnal frequency, is the mechanism by which high salinity water enters the pycnocline. [11] As discussed by Sundermeyer et al. [2005], an isolated mixing event, vertically distorting isopycnals, produces a patch of weakly stratified water that will tend to collapse laterally driven by the resulting horizontal pressure gradients. Evidence of well mixed patches may be found in individual CTD casts that show the pycnocline to have a stepped structure, with 5 10 m layers of weaker stratification 2of6

Figure 2. (a) Mean salinity (solid black line) and maximum-minimum values (dashed) from CTD casts. Mean density (s q ) in gray. (b) Horizontal qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi velocity structures for normal baroclinic modes 1 (black) and 2 (gray) computed from w n / z and normalized so that ðm 2 1 m2 p Þ ¼ 1 for the p elements in each mode n. (c e) EOF modes 1 to 3 respectively for across (black) and along-shelf (gray) currents. Figure 2c shows mode 1, accounting for 68% and 70% of the variance in the across and along-shelf currents. Similarly, 20% and 18% for mode 2 (Figure 2d), and 7% and 8% for mode 3 (Figure 2e). associated with the salinity maxima (not shown). Depending on the length and time scales of mixing, vortical modes may form, where the mixed patch starts to rotate, eventually falling into geostrophic balance [McWilliams, 1988]. The length scale to which horizontal expansion continues before being arrested by rotation will be set by the internal Rossby radius, L ro = c/f. The phase speed for long, nondispersive internal waves c =[(g H 1 H 2 )/(H 1 + H 2 )] 0.5 is calculated from an idealized 2-layer water column with surface and lower layer depths of H 1 = 30 m and H 2 = 105 m having mean densities r 1 = 1026.22 kg m 3 and r 2 = 1027.26 kg m 3 respectively (Figure 2a). For the Coriolis frequency, f, at a latitude of 49.4 N, L ro = 4.3 km. The 3 5 km wide saline lenses are therefore consistent with estimates of the internal Rossby radius suggesting that they are in near-geostrophic balance and formed by smaller scale mixing events (<5 km). 5. On-Shelf Advection and Diffusion [12] The high-salinity lenses are not restricted to the shelf edge and a net residual on-shelf flow, within the pycnocline, is required to achieve this. We will firstly estimate the advection speed by calculating the approximate rate of salt loss from a lens assumed to be propagating on-shelf over a known distance. ADCP currents will then be examined to independently support this estimate. [13] Consider a lens to be a cylinder with radius r = 2.5 km and height h = 25 m, whose salinity decreases from S 1 = 35.42 to S 2 = 35.36 over a distance of x = 50 km (Figure 1d). Assuming that this salt loss is caused primarily by vertical diffusion the rate of change of salt integrated over the volume of the cylinder is ðs 2 S 1 Þpr 2 h ¼ K zs pr 2 S K zb pr 2 S : ð1þ Dt zs zb K zs and K zb are the vertical eddy diffusivities at the top (s q = 26.4) and base (s q = 27.2) of the pycnocline respectively. Depth, z (m), is positive upwards. From vertical micro-structure measurements made near continuously between 6 8 June, the mean vertical diffusivities along the 26.4 and 27.2 potential density contours were 9.1 10 5 and 6.7 10 6 m 2 s 1 respectively. From CTD casts, S/ z s = 0.004 S m 1 and S/ z b = 0.005 S m 1 are the vertical salinity gradients at the top and base of the lens. Rearranging equation 1 to find the time taken (Dt) for the salinity of the lens to decrease from S 1 to S 2, the average speed, u, required for the lens to be advected x km in this time is u ¼ x S x K Dt ¼ zs zs K S z b zb : ð2þ hðs 2 S 1 Þ Using the above values, Dt =44daysandu =0.013ms 1. Substituting K zs and K zb with a single value of 3.9 10 5 m 2 s 1,themeanK z within the top and base of the pycnocline during the measurement period gives an estimate of 0.012 m s 1. This K z values agrees with the range of daily mean vertical diffusivities calculated on-shelf by Sharples et al. [2007] of 2 7 10 5 m 2 s 1. [14] Estimates of the horizontal diffusivity (K h ) at the scale of the lens can be made following Okubo [1971]. Based on length scales of 2 to 6 km calculated from the mean pycnocline salinity (Figure 1d), K h =0.6 2.3 m 2 s 1.Assuming that the horizontal salinity gradient, S/ x =110 5 Sm 1, is constant in all directions and taking K h = 1 m 2 s 1, the additional rate of salt loss horizontally from the lens S ( K h x 2prhSm3 s 1 ) may be included in our calculations. In doing so the estimated advection speed slightly increases to 0.018 0.020 m s 1. [15] These indirect estimates are supported by observations. Residual currents within the pycnocline, smoothed with a 12.4 hour window, for the seven days preceding the 3of6

Figure 3. Magnitude and direction of residual currents within the pycnocline (s q = 26.4 27.2) at (a) IM1 and (b) IM3 between 6 13 June 2010. Currents have been filtered with a 12.4 hour moving window to remove variability at frequencies greater than the internal tide. Note that NORTH is a northward flowing current. Scanfish transect are shown in Figure 3. At both IM1 and IM3 the magnitude of flow over a tidal cycle is 0.01 0.06 m s 1. The direction of flow between the two sites however is very different illustrating the meandering nature of residual currents over the shelf. At IM1 flow is primarily north-northwestward whereas at IM3 it is more variable and swings 270 from an initially southward to a west-northwestward direction. The time mean flows at IM1 and IM3 are 0.022 m s 1 NNW and 0.006 m s 1 WNW respectively but, given the short record and wide range in direction at IM3 they cannot be considered reliable measures of the long term net residual. Although still not a direct measure of the lenses propagation speed these observations are supportive of the estimated 0.012 0.020 m s 1 on-shelf transport. 6. Non-linear Internal Wave Transport [16] Having estimated a feasible propagation speed for the lenses, it remains to identify the mechanisms driving the residual flow. We propose that the on-shelf flux of water within the pycnocline is principally driven by non-linear mode 2 internal waves. [17] Non-linearity of the internal tide in the Celtic Sea is well recognized and is characterized by a deeply penetrating trough, more rounded wave crests and large amplitude, highfrequency non-linear waves propagating on-shelf in the trough [Pingree and New, 1995]. This asymmetry is readily observed in the thermistor chain record, as are packets of high-frequency solitons (not shown). Current variability of the internal tide is well represented by the first baroclinic mode [Pingree et al., 1986], and transport attributed to non-linearities is estimated to be of the order 1 m 2 s 1 [Huthnance, 1995]. Here we show, based on analysis of the theoretical and empirical modes of current variability, that the second baroclinic mode is more likely than the first to be driving on-shelf advection of slope water within the pycnocline. [18] The theoretical baroclinic modes, given a mean buoyancy profile N, are eigenvectors of the wave equation, 2 w/ z 2 +[(N 2 s 2 /s 2 f 2 )]k 2 w =0,wheref is the inertial frequency, k the horizontal wave number, w the vertical perturbation at depth z, and s the wave frequency. Figure 2b shows the horizontal velocity structures associated with mode 1 and 2 internal waves when s =12.4 1 hours. Dominance of the first baroclinic mode and the importance of the second is confirmed by empirical orthogonal function (EOF) analysis performed independently on the residual across and along-shelf velocities at IM3. Structures comparable to baroclinic modes 1 and 2 are captured by the first three EOFs which together account for 95% of the variance in both the across and along-shelf directions (Figures 2c 2e). [19] The first EOF (Figure 2c), containing the greatest percentage of the variance, is most comparable to the first baroclinic mode. The water column above and below 25 m oscillates in different directions, at a dominant inertial frequency. On-shelf transport in surface waters driven by nonlinearities would favor more saline surface waters and fresher bottom waters, as we observe in the mean salinity profile (Figure 2a). [20] The second two EOFs (Figures 2d and 2e), together explaining >26% of the variance, have mode 2 type structures and oscillate predominantly at the M 2 tidal frequency. Currents within the pycnocline (20 40 m) are stronger and for the third EOF in the opposite direction to those in surface and bottom waters. Both structures peak between 25 and 30 m, close to the location of the salinity maximum (30.7 m). Just as non-linear mode 1 waves sustain on average an on-shelf upper layer transport, so non-linear mode 2 current variability generates on-shelf advection within the pycnocline, and off-shelf transport in the layers above and below. Importantly, the zero crossing point for the first EOF is near the centre of the pycnocline and the mean depth of the salinity maximum. Consequently, despite the dominance of mode 1 throughout the water column as a whole, it is EOF modes 2 and 3 that make the more significant contribution to variability and residual flow within the pycnocline. [21] ThisisillustratedinFigure4wherethemeanacross and along-shelf currents between 20 and 40 m have been reconstructed. EOFs 2 and 3 combine to give a period of sustained on-shelf flow between 12:00 on 8 June and 10:00 on 10 June matching the observed time-mean flow of 0.017 m s 1 over this period (Figure 4a). The contribution from the first EOF, that has a predominately mode 1 baroclinic structure, is an order of magnitude smaller (<0.001 m s 1 ). A period of sustained along-shelf flow averaging 0.023 m s 1 is also observed between 12:00 on 10 June and 13 June (Figure 4b). The second and third EOFs contribute a mean 0.026 m s 1 current. The difference is accounted for by a weak ( 0.003 m s 1 ) southward mode 1 residual. The magnitude of 4of6

Figure 4. (a) Across and (b) along shelf currents within the pycnocline (20 40 m) smoothed with a 1 hour (dashed) and 12.4 hour (solid) running mean filter. Currents reconstructed from EOF mode 1 and modes 2 + 3 combined are in blue and red respectively. Observed currents are in black. Shaded boxes highlight time periods referred to in the text. these currents, sustained for two or more days and seemingly driven by baroclinic mode 2 type structures, fits well with the previously estimated advection speeds. Further evidence supporting the importance of mode 2 waves may be found in the thermistor chain data where the thermocline thickness oscillates with amplitudes of 1 9 m at inertial to M 2 tidal frequencies (not shown). 7. Discussion and Conclusions [22] The magnitude of tidally averaged residual flows observed within the pycnocline (0.01 0.06 m s 1 ) and the estimated advection speed (0.012 0.020 m s 1 ) compare well. Crucially however, the length of the ADCP record prevents residuals from being calculated over timescales comparable to the expected lifetime of the lenses. We are able to demonstrate that on- and off-shelf flows of the necessary magnitude are sustained for a number of days, but are unable to quantify the net residual over monthly timescales, although the observed salinity field is strong evidence that a net on-shelf flow is maintained. [23] The salt flux based estimate of propagation speed should be treated as a minimum bound. Firstly, we assumed that the lenses originated from the nearest shelf edge point to the transect and were advected along it. The wide range of flow directions observed however demonstrate that the path of each individual lens is unlikely to be so direct. Also, estimates of K z were made over a neap tide; greater diffusivities and faster advection speeds would be expected during springs. A number of processes are likely to contribute to a meandering on-shelf flow and patchy generation of saline lenses. Firstly, the shelf edge is not smooth; rough and variably sloping topography mean that increased mixing does not take place in one long band along the shelf edge. Secondly, along-slope as well as across-slope flow is responsible for the generation of high frequency wave packets and diapycnal mixing [Holt and Thorpe, 1997]. In this way, internal waves propagate on-, off- and along-shelf dependent upon their generation mechanism. [24] The estimated propagation speed is dependent upon the diffusivity, which may vary by two orders of magnitude. Minimum values observed were within the range of molecular diffusivity (4 7 10 7 m 2 s 1 ). Maximum values at the top and bottom of the pycnocline were 8.1 10 4 m 2 s 1 and 8.5 10 5 m 2 s 1 respectively. Using these maxima, the estimated propagation speed of 0.12 m s 1 remains realistic (50 km in 4-5 days). These speeds are observed, but maintained for hours rather than days (Figure 4). Given the patchy and wide ranging nature of diffusivity values it was appropriate to base an estimate of the propagation speed of each lens on the average diffusivity that it experiences. [25] The strength of the pycnocline may be an important controlling factor in the persistence of these features. A threelayer water column with a broad pycnocline, as seen here, as opposed to a tighter two-layer structure, would better support higher mode internal waves and thus advection by mode 2 baroclinic currents. Additionally, within a wide pycnocline, where the diapycnal diffusivity is generally low, straining from oscillatory mode 1 current shear is insufficient to cause significant horizontal dispersion of the lenses [Young et al., 1982]. [26] The importance of this new mechanism that can transport slope water hundreds of kilometers onto the continental shelf is yet to be explored from a biogeochemical perspective. Increased mixing at the shelf edge drives a vertical flux of inorganic nitrate into the base of the pycnocline [Sharples et al., 2007]. Although this nitrate is used rapidly, the organic matter produced may be transported onshelf within the lenses and ultimately be recycled. [27] Acknowledgments. This work was funded by UK NERC grants, FASTNEt (NE/I030224/1) and Inertial Mixing (NE/F002432/1). ADCP and micro-structure profiler data were provided by Chris Old and Yueng-Djern Lenn (Bangor University, Wales). Thanks to Jeff Polton for helpful discussion. AVHRR SST data were supplied courtesy of NEODAAS, Plymouth. [28] The Editor thanks two anonymous reviewers for their assistance in evaluating this paper. References Holt, J. T., and S. A. Thorpe (1997), The propagation of high frequency internal waves in the Celtic Sea, Deep Sea Res., Part I, 44(12), 2087 2116. Huthnance, J. M. (1995), Circulation, exchange and water masses at the ocean margin: The role of physical processes at the shelf edge, Prog. Oceanogr., 35, 353 431. Huthnance, J. M., J. T. Holt, and S. L. Wakelin (2009), Deep ocean exchange with west-european shelf seas, Ocean Sci., 5, 621 634. Inall, M. E., D. Aleynik, T. Boyd, M. Palmer, and J. Sharples (2011), Internal tide coherence and decay over a wide shelf sea, Geophys. Res. Lett., 38, L23607, doi:10.1029/2011gl049943. Lentz, S. J. (2003), A climatology of salty intrusions over the continental shelf from Georges Bank to Cape Hatteras, J. Geophys. Res., 108(C10), 3326, doi:10.1029/2003jc001859. Liu, K. K., L. Atkinson, R. A. Quiñones, and L. Talaue-McManus (2010), Biogeochemistry of continental margins in a global context, in Carbon and Nutrient Fluxes in Continental Margins: A Global Synthesis, edited by K. K. Liu et al., pp. 3 24, Springer, Berlin, Germany. McCave, I. N., I. R. Hall, A. N. Antia, L. Chou, F. Dehairs, R. S. Lampitt, L. Thomsen, T. C. E. van Weering, and R. Wollast (2001), Distribution, composition and flux of particulate material over the European margin at 47 50 N: Results from OMEX I, Deep Sea Res., Part II, 48, 3127 3139. McWilliams, J. C. (1988), Vortex generation through balanced adjustment, J. Phys. Oceanogr., 18, 1178 1192. New, A. L., and R. D. Pingree (1990), Large amplitude internal soliton packets in the central Bay of Biscay, Deep Sea Res., Part A, 37, 513 524. Okubo, A. (1971), Oceanic diffusion diagrams, Deep Sea Res. Oceanogr. Abstr., 18, 789 802. 5of6

Pingree, R. D., and A. L. New (1995), Structure, seasonal development and sunglint spatial coherence of the internal tide on the Celtic and Armorican shelves and in the Bay of Biscay, Deep Sea Res., Part I, 42(2), 245 284. Pingree, R. D., G. T. Mardell, and A. L. New (1986), Propagation of internal tides from the upper slopes of the Bay of Biscay, Nature, 321, 154 158. Rippeth, T. P., and M. E. Inall (2002), Observations of the internal tide and associated mixing across the Malin Shelf, J. Geophys. Res., 107(C4), 3028, doi:10.1029/2000jc000761. Sharples, J., et al. (2007), Spring-neap modulation of internal tide mixing and vertical nitrate fluxes at a shelf edge in summer, Limnol. Oceanogr., 52(5), 1735 1747. Sundermeyer, M. A., J. R. Ledwell, N. S. Oakey, and J. W. Greenwood (2005), Stirring by small-scale vortices caused by patchy mixing, J. Phys. Oceanogr., 35, 1245 1262. Young, W. R., P. B. Rhines, and C. J. R. Garrett (1982), Shear-flow dispersion, internal waves and horizontal mixing in the ocean, J. Phys. Oceanogr., 12(6), 515 527. 6of6