<<

VOLUME 47 JOURNAL OF PHYSICAL OCEANOGRAPHY JUNE 2017

Topographic and Mixed Layer Submesoscale Currents in the Near-Surface Southwestern Tropical Pacific

KAUSHIK SRINIVASAN,JAMES C. MCWILLIAMS, AND LIONEL RENAULT Department of Atmospheric and Oceanic Sciences, University of California, Los Angeles, Los Angeles, California

HRISTINA G. HRISTOVA Joint Institute for Marine and Atmospheric Research, University of Hawaii, Honolulu, Hawaii, and NOAA/Pacific Marine Environmental Laboratory, Seattle, Washington

JEROEN MOLEMAKER Department of Atmospheric and Oceanic Sciences, University of California, Los Angeles, Los Angeles, California

WILLIAM S. KESSLER NOAA/Pacific Marine Environmental Laboratory, Seattle, Washington

(Manuscript received 15 September 2016, in final form 28 January 2017)

ABSTRACT

The distribution and strength of submesoscale (SM) surface layer fronts and filaments generated through mixed layer baroclinic energy conversion and submesoscale coherent vortices (SCVs) generated by topo- graphic drag are analyzed in numerical simulations of the near-surface southwestern Pacific, north of 168S. In the Coral a strong seasonal cycle in the surface heat flux leads to a winter SM ‘‘soup’’ consisting of baroclinic mixed layer eddies (MLEs), fronts, and filaments similar to those seen in other farther away from the equator. However, a strong wind stress seasonal cycle, largely in sync with the surface heat flux cycle, is also a source of SM processes. SM restratification fluxes show distinctive signatures corresponding to both surface cooling and wind stress. The winter peak in SM activity in the is not in phase with the summer dominance of the mesoscale eddy kinetic energy in the , implying that local surface layer forcing effects are more important for SM generation than the nonlocal eddy deformation field. In the to- pographically complex Solomon and Bismarck , a combination of equatorial proximity and boundary drag generates SCVs with large-vorticity Rossby numbers (Ro ; 10). River outflows in the Bismarck and Solomon Seas make a contribution to SM generation, although they are considerably weaker than the to- pographic effects. Mean to eddy kinetic energy conversions implicate barotropic instability in SM topographic wakes, with the strongest values seen north of the along the coast of Papua .

1. Introduction vertical and horizontal extent of these eddies scale with the thermocline depth and the thermocline baroclinic Mesoscale eddies are prominent throughout the deformation radius R , respectively (Tulloch et al. global , with kinetic energies (KEs) that are typi- d 2011). Generated through baroclinic instability in the cally about an order of magnitude larger than the mean thermocline, and with typical midlatitude horizontal kinetic energy (Stammer 1997; Chelton et al. 2011). The scales of around 50 km, these eddies have small Rossby (Ro 5 U/fL 1) and Froude numbers (Fr 5 U/HN 1) NOAA/Pacific Marine Environmental Laboratory Contribution and are correspondingly geostrophically balanced. Number 4569 and Joint Institute for Marine and Atmospheric Here, U, L, and H are velocity, length, and depth scales Research Contribution Number 16-394. of the mesoscale eddies, while f and N are the Coriolis parameter and associated buoyancy frequency. In cer- Corresponding author e-mail: Kaushik Srinivasan, kaushiks@ tain regions, like the Coral Sea, mesoscale eddies can atmos.ucla.edu also be generated by barotropic instability of the nearly

DOI: 10.1175/JPO-D-16-0216.1 Ó 2017 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses). 1221 Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1222 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47 zonal jets in the region (Qiu et al. 2009). Ageostrophic cyclostrophic force), as opposed to the usually geo- motions become significant and dynamically relevant at strophically balanced mesoscale eddies. This is as much a larger Rossby numbers (i.e., smaller length scales) and dynamical definition of SM as it is one of scale (McWilliams are often referred to as submesoscale currents or more 1985; Molemaker et al. 2015; Gula et al. 2015). concisely as submesoscales (SMs). A variety of routes to In this paper, we examine SM statistics and dynamics the generation of SM features have been identified in in the geographically diverse southwestern Pacific re- recent papers and typically occur at the ocean surface gion north of 168S(Fig. 1). The surface waters in the and bottom (Molemaker et al. 2015). Mixed layer in- region are largely supplied by the westward flowing stability (MLI; Boccaletti et al. 2007) of the weakly South Equatorial Current (SEC), which bifurcates at the stratified surface mixed layer leads to the formation of coast of into northward and southward west- mixed layer eddies (MLEs) and fronts (Fox-Kemper ern boundary currents. Somewhat uniquely, the north- et al. 2008; Capet et al. 2008a) with horizontal scales ward western boundary current, the Gulf of Papua ;10 km or less. MLEs are stronger for larger horizontal Current, is strongest subsurface (below 300 m) and re- buoyancy gradients =hb and deeper mixed layers Hb and mains so as it travels into the Solomon and Bismarck are an important reason for the significant enhancement Seas. Before its bifurcation, the SEC waters form a se- in SM activity during winter. Frontogenesis, originally ries of alternating zonal jets in the Coral Sea by inter- identified in the context of the atmosphere (Hoskins and acting with the islands chains of New Caledonia, Bretherton 1972), occurs when the mesoscale eddy Vanatu, and Fiji (Couvelard et al. 2008). Complex strain field enhances =hb in the mixed layer through the coastlines, island chains, and subsurface topographic action of a dynamically active ageostrophic secondary features result in a high variability of the eddy kinetic circulation, leading to the formation of fronts that can be energy (EKE) over seasonal (Hristova et al. 2014) and clearly seen at the ocean surface in observations and annual time scales (Kessler and Cravatte 2013; Melet models. While frontogenesis can occur independently et al. 2010). We show in this study that the interactions of from MLI, the two processes often accompany each the surface currents with topography are a significant other with secondary frontogenesis arising on the edges source of SCVs in the Solomon and Bismarck Seas. Our of MLEs. Furthermore, unlike the preceding inviscid focus is on SMs in the near-surface ocean (top 100 m). mechanisms of SM generation and maintenance, a Many recent papers have examined the southwestern nonconservative SM process is associated with the sec- Pacific using both observational and modeling ap- ondary circulation and frontogenesis due to turbulent proaches. Studies using mesoscale-resolving models and vertical momentum mixing in fronts and filaments (Gula Argo and altimetry datasets have focused on seasonal et al. 2014; McWilliams et al. 2015), referred to as the and interannual variability in the (Kessler turbulent thermal wind (TTW) balance. In general, all and Cravatte 2013; Melet et al. 2010). These studies three dynamical processes are expected to be active in the show that the upper ocean has an ENSO-like variability, surface layer, although their relative importance depends since the Solomon Sea lies within the equatorial warm on the nature and strength of the (local) surface forcing pool. Mesoscale eddies are the dominant variability on and the (nonlocal) mesoscale eddy strain field. One of the seasonal and subseasonal time scales, with anticyclones aims of the present study is to examine the spatiotemporal showing a weak dominance compared to the cyclones evolution of mixed layer restratification fluxes and their (Hristova et al. 2014). The maximum EKE in the Solo- dynamical attribution to the three SM processes discussed mon Sea is distributed over all of its central region in above in the southwestern tropical Pacific. agreement with satellite altimetry data (Hristova et al. A different route to generating eddies, both meso- 2014; Gourdeau et al. 2014). Djath et al. (2014) uses a scale and SM, is the interaction of mean and mesoscale submesoscale-permitting model of the Solomon Sea on a currents with topography. The basic mechanism horizontal grid resolution of 1/368 (around Dx 5 3km) [explained in Molemaker et al. (2015)] involves the to study spectral properties of velocity in the Solomon boundary separation of high-vorticity sheets generated Sea. They find a progressive shallowing of energy by bottom drag in the bottom boundary layers and wavenumber spectral slopes with increasing resolution subsequent barotropic or centrifugal instability that and the possibility of a forward kinetic energy cascade generates SM vortices and turbulence. The separated at small scales. In this study, we characterize the spatial wakes have a ;0.1–10-km horizontal scale and roll up structure, seasonal variability, and dynamics of SMs in into either mesoscale or SM coherent vortices (SCVs), the tropical southwestern Pacific region in and around depending on whether the vortex Rossby number Ro 5 the Solomon Sea. We use a hierarchy of nested, re- z/f (where the vorticity z 5 ›xy 2 ›yu) is small or O(1). alistic geometry cases in the Regional Oceanic Mod- SCVs are gradient wind balanced (i.e., including the eling System (ROMS), with horizontal resolutions as

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1223

2 23 FIG. 1. Maps of the annually averaged surface fluxes (m s ) of (a) heat, (b) freshwater, (c) buoyancy [with 2 annually averaged wind stress (N m 2) overlaid], and (d) the bottom topography in the region. Four analysis sub- domains are indicated with black lines in (b): Bismarck Sea, Solomon Sea, Gulf of Papua, and Coral Sea.

fine as Dx 5 500 m, to analyze SM behavior in several s-coordinate oceanic model (Shchepetkin and McWilliams distinctive of the southwestern Pacific. The 2003, 2005). Momentum advection is computed using a paperisorganizedasfollows:Section 2 describes the third-order, upwind-biased scheme (Shchepetkin and model details and numerical methodology. Section 3 McWilliams 2005) that is the equivalent to a fourth- characterizes the statistics of mesoscales and sub- order central scheme supplemented by a biharmonic mesoscales in this region and their seasonal variability. diffusion operator whose local diffusivity depends In section 4, the SM dynamics is analyzed by looking at the both on velocity and grid size Dx, such that it vanishes spatiotemporal variability and vertical structure of the as Dx / 0. Tracer advection is performed using an available potential energy (APE) conversion to SM kinetic isoneutral advection scheme that helps preserve water energy (i.e., the restratification flux). This section also mass properties (Lemarié et al. 2012). presents a comparison of this model-based restratification a. Model cases flux with existing theories and parameterizations for SM turbulence in the mixed layer. Section 5 examines the en- The simulations are computed using a set of nested ergetics of topography-induced submesoscale generation ROMS cases starting from a case with Dx 5 12 km that at the island chains and straits in and around the Solomon spans a decade in time and the entire Pacific Ocean. The and Bismarck Seas, while the final section summarizes and boundary and initial conditions for this run are gener- concludes this manuscript. ated from SODA climatology (Carton and Giese 2008). A nested Dx 5 4 km case is then computed for the region between the latitudes of 408S and 108N and the longi- 2. Methodology tudes 1308E and 1608W. The boundary and initial con- All the simulations are made using ROMS, a primitive ditions are constructed from the 12-km output fields equation, split–explicit, hydrostatic, terrain-following, using the methodology explained in Mason et al. (2010).

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1224 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

The Dx 5 4 km case has also been the subject of a recent the atmospheric forcing, but intraseasonal forcing is al- analysis of the mesoscale structure and variability in this lowed. The surface heat fluxes are obtained from the 18 region (Hristova et al. 2014). Employing a similar Common Ocean–Ice Experiment (CORE) monthly methodology, a Dx 5 1.5 km case is computed for a climatology over the period 1981–2006, supplemented period spanning a year, and it can be considered to be with a weak restoring toward a Pathfinder SST clima- ‘‘submesoscale permitting’’ in the sense that at least the tology. The freshwater fluxes are obtained from Ham- larger SM eddies and fronts are resolved. The spinup burg Ocean Atmosphere Parameters and Fluxes from time in both cases is just over 3 months, and the model Satellite Data (HOAPS-3) climatology over the period data during that period are excluded from the analyses 1987–2005 at 1/28 resolution with a weak restoring toward presented here. The vertical grid structure in all three the World Ocean Atlas 2005 climatology. The river cases of the hierarchy is identical and consists of N 5 runoff is simulated as a surface precipitation, spatially 51s levels. Our highest-resolution run, a Dx 5 500 m case, spread around the river mouth. The annual-averaged nested down from Dx 5 1.5 km spans the winter months heat and salinity fluxes are shown in Fig. 1, expressed as of June, July, and August and has N 5 100s levels. The contribution to the surface buoyancy flux Bo,where 0 Dx 5 500 m run is, however, only used marginally in the buoyancy is rescaled density b 52gr /r0. The flux Bo is present paper and will be a subject of future studies, related to the surface heat Q and freshwater with a majority of the results here based on the Dx 5 [evaporation 2 precipitation (E 2 P)] fluxes by 1.5 km run. The vertical grid stretching is done with the parameters u 5 6, u 5 1.5, and h 5 250 m [for an s b c B 5 gaQ /(r C ) 2 gb(E 2 P)S . (1) explanation of the ROMS vertical grid structure, see o |fflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflffl}o 0 p |fflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflffl} 2 Shchepetkin and McWilliams (2009)]. The ocean BT BS surface and bottom are parameterized using the K-profile parameterization (Large et al. 1994), which In Fig. 1b, there is a net freshwater flux north of 128S has seen widespread use in ocean and climate models and in the river mouths north and south of Papua New (Danabasoglu et al. 2006). Its precise formulation in Guinea. The seasonal winter cooling in the Coral Sea is ROMS differs from Large et al. (1994) in determining the also prominently visible. Even though the annual av- boundary layer depth hBL with an integral functional erages seem to imply comparable contributions from composed of the stratification, shear, rotation, and tur- the heat flux and the net freshwater flux, we see in Fig. 2 bulent buoyancy flux influences (McWilliams and Huckle that the variability in the BT is up to a factor of 3 greater 2006; McWilliams et al. 2009; Lemarié et al. 2012). This than in BS, indicating that the intraseasonal variations approach, explained in detail in Lemarié et al. (2012), might be dominated by temperature, a fact that is alleviates some of the computational oscillations poten- confirmed in section 4b. The Gulf of Papua has the tially present, leading to improved spatial smoothness. weakest winter cooling signal, with strong surface heating and freshwater outflows at the river mouths. b. Surface wind stress and buoyancy forcing The spatially averaged BS seasonal cycle in Fig. 2 in- The surface wind stress and heat and freshwater fluxes cludes contributions from both riverine outflows and are constructed using the approach detailed in Lemarié surface (E 2 P) forcing. While the total annual fresh- et al. (2012). This is done with a two-step procedure. The water input in the Solomon Sea is largely due to the first step is used to generate a wind stress climatology, surface precipitation term P in the Bismarck Sea and and the second step is to add its daily variability. The Gulf of Papua regions, river outflows have significant climatology is the QuikSCAT-based Scatterometer contributions. The total annual riverine output into the Climatology of Ocean Winds (SCOW) for the years Gulf of Papua is comparable to other river systems; in 2004–09, corrected using buoy data from the Tropical particular, it is about half of the Mississippi river out- Atmosphere–Ocean Array in the equatorial Pacific. The flow. Surface precipitation in the Solomon and Bis- daily variability is obtained from the Centre ERS marck subdomains account for most of the BS with a d’Archivage et de Traitement (CERSAT; French ERS smaller fraction accounted for by the riverine outflows. processing and archiving facility) wind anomalies for the The seasonal variability of the wind stress, t and its period July 2005 to June 2006 with an adjustment to standard deviation are also shown in Fig. 2.Themean ensure periodicity in time for this interval. Thus, a single, wind stress varies contemporaneously with the surface year-long, periodic, daily wind product is obtained and heat flux, reflecting the atmospheric Hadley circulation. then repeated to generate multiyear forcing. The pres- In general, within the spatial extent of the Dx 5 1.5 km ent simulations are therefore idealized process studies, case, the mean surface wind stress increases with in- in the sense that interannual variability is excluded from creasing latitude.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1225

2 23 FIG. 2. (top) Seasonal cycle of spatially averaged surface fluxes of heat, freshwater, and buoyancy (m s ). (bottom) Seasonal cycle of xy 2 t xy t02 1/2 2 spatially averaged surface wind stress h im and its standard deviation h im (N m ). The spatial averaging is performed over the four regions delineated in Fig. 1.

The choice of the spatial averaging subdomains in we analyze the spatiotemporal statistics of the surface Fig. 1 is somewhat arbitrary. The Coral Sea is a simple vorticity, horizontal velocity divergence, buoyancy gra- choice, bereft of near-surface topographic effects or dient, and frontogenetic tendency, quantities that are freshwater fluxes. A prominent seasonal cycle in wind associated with SM fronts and filaments (Capet et al. and buoyancy fluxes allows a careful examination of 2008b). The EKE, which is dominated by the mesoscale mixed layer SMs. The Gulf of Papua subdomain is eddy field and is the subject of numerous previous chosen to study the role of freshwater forcing and as- studies, is also briefly discussed. sociated fronts. The Solomon and Bismarck subdomains a. Eddy kinetic energy examine the effects of equatorial proximity on SM cur- rents. In section 5, a different set of near-boundary av- The annual cycle and dynamics of mesoscale eddies eraging regions is chosen to specifically examine the has been previously analyzed by Hristova et al. (2014) relative influence of topographic and mixed layer gen- using the same Dx 5 4 km model case used here. It shows eration of SMs. good agreement between the geostrophic EKE in the The spatiotemporal averages in this paper are the model compared to satellite altimetry measurements. following: Temporal averaging is performed over Here, we briefly compare the EKE from the previously monthly hi , seasonal hi , or annual time scales hi . analyzed 4-km case with the Dx 5 1.5 km case. The EKE m s xy a Spatial averaging operators are written as () , with the here is computed from velocity anomalies derived rel- region of averaging typically being one of the four ative to a climatological monthly mean. In Fig. 4, we see xyz choices in Fig. 1, while volume averages () include that the mesoscale variability in the Solomon Sea dis- the depth dimension with the range between the surface plays similar spatial patterns and magnitudes, in spite of and the base of the mixed layer. the shorter time of averaging (hence greater estimation error) in the Dx 5 1.5 km case, implying a degree of convergence of EKE estimate with model resolution. 3. Mesoscale and submesoscale statistics This is because the mesoscale eddies and boundary Figure 3 shows a snapshot of surface vorticity based on current fluctuations dominate the EKE, unlike most of simulations collated from the three nested cases in the the other quantities examined in the subsequent sub- winter season (in this paper all sections that are dominated by SM variability. Some seasons are relative to the Southern Hemisphere). The differences do arise in the spatial patterns of EKE in the surface vorticity patterns of the 4-km case are domi- Coral Sea, however, although the magnitudes are of a nated by eddies, but higher-resolution cases also prom- similar order. In fact, EKE values averaged over each inently display SM frontal patterns. A distinct pattern of year of the Dx 5 4 km case in the Coral Sea have dif- reduced magnitude of vorticity is seen on approaching ferences with the same order of magnitude and spatial the equator (where geostrophic balance is weaker). Next structure as seen between the Dx 5 4 and 1.5 km cases

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1226 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

21 FIG. 3. A snapshot of the magnitude of surface vertical vorticity z (s ) from the nested ROMS hierarchy of Dx 5 4, 1.5, and 0.5 km cases. Their respective domains are outlined by black boxes. The date of the snapshot is 1 Jun.

(not shown). This is not a serious impediment to our 148–188S as caused by barotropic instability of the al- study of SM currents, which are much smaller than ternating zonal current system (the North Vanatu jet– mesoscale eddies and have much shorter lifetimes Coral Sea Countercurrent–New Caledonia jet Current (typically a few days compared to a few weeks for me- System) in this region. Using a reduced gravity shallow soscale eddies), thereby allowing more accurate SM water model that compares well with the results from statistical estimations. satellite altimetry, they find an EKE peak in the months In a previous study of the Coral Sea mesoscales, of December–January (DJ) and a minimum in May– Qiu et al. (2009) attributes the EKE in a latitudinal band June (MJ). Our analysis of the EKE seasonal cycle in

2 22 FIG. 4. Spatial patterns of the annually averaged surface EKE (m s ) in and around the Solomon Sea from the Dx 5 (a) 4 and (b) 1.5 km cases.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1227 this latitude band leads to similar results (not shown). In in fronts and filaments similarly exhibits a secondary sum, the EKE variation in the region is out of phase with circulation that leads to cyclonic and downwelling the seasonal local surface forcing of wind and buoyancy dominance (McWilliams et al. 2015; Gula et al. 2014). flux (Fig. 2), raising the intriguing question of whether Thus, the near-surface w, the surface horizontal di- the nonlocal forcing (the mesoscale eddy deformation vergence d 5 ›xu 1 ›yy, and z are expected to have finite field) or the local surface forcing is more important for skewness in regions of active SM frontal dynamics. SM generation. Below we show that local forcing effects We compute temporal statistics for the model cases typically dominate, mirroring previous studies of mid- (with instantaneous output files available twice a day) at latitude SM seasonal cycles (e.g., Mensa et al. 2013) that each point in our domain. Spatial maps of these tem- lack the strong mesoscale seasonal cycle observed in poral statistics for the summer [January–March (JFM)] this region. and winter [July–September (JAS)] season are pre- sented in Figs. 5 and 6. The skewness statistic is masked, b. Vertical velocity and vorticity subject to a variance constraint skew[hd02i , C] 5 0, 2 2 The quasigeostrophic (QG) equations have a funda- where C 5 3 3 10 11 s 2, with a similar criterion used to mental cyclone–anticyclone symmetry (z 4 2z) that is mask the vorticity skewness. This focuses attention on approximately validated in observed (Chelton et al. regions with appreciable SM variability. 2011) and modeled (Tulloch et al. 2011) analyses of 1) CORAL SEA SUBMESOSCALE ‘‘SOUP’’ mesoscale eddies. QG mesoscale eddies also have a relatively weak vertical velocity w, solely because of the The largest seasonal contrast is in the statistics for deflection of isopycnals over the spatial scale of the both z and d in the Coral Sea. This is indicative of a pycnocline baroclinic deformation radius Rd (’50 km in strong SM activity in the winter (Fig. 3), with the large the midlatitudes and even larger in the tropics). Recent skewness in z and d demonstrating clear cyclonic and studies (Callies et al. 2015; Chavanne and Klein 2016) downwelling dominance. Similar seasonal cycles have using mixed layer QG models predict the emergence of been found in midlatitude simulations (Mensa et al.

MLEs with a size much smaller than Rd, associated with 2013; Gula et al. 2014) and observations (Callies et al. the somewhat less clearly defined (owing to ambiguities 2015). Previous studies have largely interpreted the in defining the surface layer stratification)pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi mixed layer seasonal SM as being a consequence of MLEs, but we deformation radius Rml [Rml 5 HbDbml/f ’ O(1) km] show in section 3d that frontogenesis and TTW pro- in midlatitudes. The associated vertical velocity is cor- cesses are also consistent with this seasonal cycle. The respondingly much larger than in thermocline QG primary reason for this SM seasonality is that the surface models, but in QG models it has an upwelling– wind stress and surface heat flux, which control the downwelling (w 4 2w) symmetry concomitant with boundary layer mixing and Hb, also have a strong sea- the z 4 2z symmetry. The QG dynamics is formally sonal cycle (Fig. 2). The Coral Sea panels of Fig. 7 un- invalid when Ro ; 1, which can happen when the strain derscore these results: the statistical indicators of SM field generated by mesoscale and MLEs rapidly in- activity are clearly stronger in the Dx 5 1.5 km case. creases the existing =hb to large values (provided =hb has a Perhaps more interestingly, there is small temporal nonzero projection on the direction of local principal phase difference in the peak values of both d and strain), leading to the formation of fronts and filaments. z variance between the cases: the Dx 5 1.5 km case has One of the dynamical mechanisms for frontogenesis its peak activity in July–August as opposed to Septem- involves an O(Ro) ageostrophic secondary circulation ber for the 4-km case. in the cross-frontal plane generated by the strain field, 2) BISMARCK AND SOLOMON SEAS TOPOGRAPHIC which has an upwelling, anticyclonic, lighter side and a SUBMESOSCALES downwelling, cyclonic, heavier side. On the two sides, cyclonic vorticity is enhanced and anticyclonic vorticity In contrast to the open Coral Sea, large d and is diminished by the vortex stretching and compression z variance values are seen along the eastern coasts of in the vorticity relation and (Fig. 5) in the west Solomon Sea off the Woodlark Archipelago and gen- Dz ›w erally along island wakes visible as coastally spun eddies ’ (f 1 z) . (2) Dt ›z and fronts in the snapshots in Figs. 3 and 8. These are a result of the separation of bottom drag–generated shear In addition, because of the finite Ro asymmetry, the layers in shallow water that spawn near-surface SM downwelling w on the cyclonic side is larger in a front eddies. They are especially prevalent near coastal than the upwelling on the lighter side. The TTW balance headlands and islands with strong passing currents, for

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1228 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

21 FIG. 5. Spatial patterns of the rms variability of surface horizontal divergence d0 and vertical vorticity z0 (s ) and their associated skewness computed over the summer months of January, February, and March in and around the Solomon Sea from the Dx 5 1.5 km case. The skewness calculation is spatially masked (white space) using a vari- ance threshold to avoid regions of weak variability. example, the northward flow during the winter through As we show in section 3c, freshwater-induced fronts do the relatively Vitiaz Strait and the SEC along the eastern occur in this region but are strongest between the months coast of New Ireland. The equatorial proximity and of February to May, which do not have significant local small island sizes off the Vitiaz Strait and New Ireland maxima in the right panels of Fig. 7. ensure that the eddy generation involves SM dynamics A higher-resolution (Dx 5 500 m) Solomon Sea case with rather large Rossby numbers ;10 (Dong and during the winter season (Fig. 9) has, as expected, a McWilliams 2007). much stronger SM signal relative to the Dx 5 1.5 km The skew(z) values in the winter island wakes impli- case. As with the Dx 5 1.5 km case, the Dx 5 500 m SM cate anticyclonic generation (red patches in Fig. 5), features are dominated by eddies and fronts generated consistent with the sign of vorticity generated by through interaction of mesoscale flow with the topo- northwestward currents flowing along these coasts. graphic features off the Woodlark Archipelago and the Similar anticyclonic SCVs have been studied in detail by Solomon Islands to the east (not shown). What is sig- Molemaker et al. (2015) off the California coast due to nificant, however, is a strong signature of secondary California Undercurrent separation off a headland. The frontal instabilities and internal waves in the Dx 5 500 m summer months also show smaller but significant SCVs case that are poorly resolved in the Dx 5 1.5 km case in the Bismarck Sea and off the eastern Papua New (not shown). These features are notably absent in mid- Guinea coast and in some of the island wakes sur- latitude simulations of comparable resolution (e.g., rounding the Solomon Sea. These are principally to- Gula et al. 2014). One reason for this is that the larger pography related (boundary generated and consistent baroclinic deformation radius (’200 km) and the mixed with the wind stress seasonal cycle), but off the eastern layer deformation radius (’5 km) in the Solomon Sea, Papua New Guinea coast itself, some generation can be compared to the midlatitudes at the same grid resolu- attributed to freshwater-generated fronts off the river tion, has higher effective resolution, owing to equatorial mouths. Such fronts were recently shown to be of dy- proximity. However, because this paper is a broad study namical importance off the coast of the (more voluminous) of SMs in this region, a detailed dynamical description of Mississippi River in the (Luo et al. 2016). the Dx 5 500 m case is deferred to future studies, and the

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1229

FIG.6.AsinFig. 5, but for the months of June, July, and August.

SM statistics and dynamics shown in this paper are November (the choice of November is arbitrary with largely derived from the Dx 5 1.5 km case. similar eddies seen all year; Fig. 10; Rd ’ 200 km at this latitude). Three large and persistent mesoscale eddies 3) THE EAST SOLOMON SEA SUBMESOSCALE are seen during this month, with weak values of buoy- ‘‘DESERT’’ ancy gradient and frontogenetic tendency in the eddy A persistent feature of the statistical maps in Figs. 5 centers. This is consistent with the results in Capet et al. and 13 (below) is the weakness of SM signatures in the (2008a,b) that eddy centers typically have strong vor- eastern Solomon Sea, year-round. This is in spite of the ticity values and weak values of strain and j=hbj, while large EKE signal (Fig. 4) associated with the presence of the opposite happens near the eddy boundaries where large mesoscale eddies in the region, seen, for example, SM processes are most active. Thus, the mesoscale strain in maps of the vorticity averaged over the month of field is conducive to fronts and frontogenesis at the eddy

22 FIG. 7. Seasonal cycle of surface d and z variance (s ) averaged over the four regions for the Dx 5 4 (black) and 1.5 km (blue) cases. Notice the strong increase in these SM indicators with resolution. See Fig. 9 forcomparisonwiththeDx 5 500 m case during the winter months.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1230 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

22 FIG. 8. Snapshots of surface z/f, d/jfj, and j=hbj (s ) in the Bismarck Sea on 1 Jul from the Dx 5 1.5 km case. Note that July is the month of peak SM variance in this region (see Fig. 7). edges. While these maps only depict the month of month of June by examining zonally averaged statistics November, similar large eddies are usually prevalent of previously discussed SM proxies. To allow for a in the eastern Solomon Sea and give rise to the SM relative comparison of their latitudinal variability, we desert seen in the statistical maps of Figs. 5 and 13.An normalize each zonally averaged statistic with their alternative approach to characterizing the relative in- maximum values over the domain. [Thus, we are = x = x fluences of vorticity and strain is through the Okubo– plotting hj hbj im/max(hj hbj im) and so on.] Then by 2 2 2 2 Weiss parameter Q 5 z 2 S ,whereS 5 (ux 2 yy) 1 selectively masking out the near-boundary regions 2 (uy 1 yx) is the trace of the strain tensor. In a (i.e., within 90 km of the boundary) before zonal av- submesoscale-permitting study in the tropical North eraging (Fig. 11a) and comparing with the unmasked Atlantic, Veneziani et al. (2014) find that the equatorial zonal averages (Fig. 11b), we find that in this region the regions are dominated by strain Q , 0, similar to the SMs in the near-equatorial regions are primarily due to situationintheSolomonSea,resultinginweaker topographic effects. Thus, north of about 68S, the SM frontogenesis. generation mechanism can be thought to transition from mixed layer to topographic on approaching the 4) LATITUDINAL VARIATION OF SUBMESOSCALE equatorial region. STATISTICS c. Frontogenetic tendency In Fig. 3, a general decrease of both mesoscale eddies and SM fronts, away from the boundaries, was Fronts are formed when horizontal density gradients seen on approaching the equator. This is qualitatively are enhanced by various dynamical processes that in- expected because the mechanisms of baroclinic in- clude mesoscale strain, MLEs, and TTW-induced sec- stability and frontogenesis both require rotation. A ondary circulations (section 1). From the buoyancy quantitative view of the decrease of SM amplitudes equation, the Lagrangian frontogenetic tendency can be as a function of latitude is displayed in Fig. 11 for the written as

2 22 2 22 25 FIG. 9. Seasonal cycle of monthly mean values of surface d0 (s ), z0 (s ), F (s ), and 22 =hb (s ) averaged over the Solomon Sea area for the Dx 5 1.5 km (blue) and 500 m (red) cases. The Dx 5 500 m case is only available during the winter months.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1231

25 22 FIG. 10. Maps of surface hzim/f, hF im (s ; section 3c), and hj=hbjim (s ) in the Solomon Sea, averaged over the month of November for the Dx 5 1.5 km case.

Dj= bj2 frontogenetic (F . 0), and this component is defined as h 5 F , (3) F 5 = Dt Q hb, where F 52 › › 1 › y› › › 1 › y› where is the frontogenetic tendency consisting of Q ( xu xb x yb, yu xb y yb). (4) advective, vertical mixing, and horizontally diffusive terms (Capet et al. 2008b). The diffusive and mixing We construct temporal averages of j=hbj and F in a terms are generally frontolytic (i.e., F , 0), although manner similar to the d and z statistics in section 3b.In the latter has a more complex effect for the case of TTW both these quantities the SM components dominate. The fronts and filaments (Gula et al. 2014). Fronts and fila- dominance of frontogenesis over frontolysis is evident in ments are typically formed when the advective part is the probability distribution function (PDF) of surface F

FIG. 11. Latitudinal variation of normalized [see section 3b(4) text for details], zonally av- eraged quantities averaged over the month of June at the surface. (a) The mesoscale filtered x x 22 fields for buoyancy gradient hj=hbMEj i and the total field hj=hbj i (s ) and the analogous x m m x 2 z02 02 2 vorticity quantities h im and hd im (s ), where regions within 90 km of the boundary are masked out before zonal averaging. (b) As in (a), but with zonal averaging performed over the entire domain without masking the boundary regions. These are for the Dx 5 1.5 km case.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1232 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

The regional seasonal cycles of the vertical structure of hF i in Fig. 14 are somewhat challenging to interpret. The Coral Sea cycle shows a winter enhancement, as

expected, from the deeper Hb values associated with the effects of wind stress and surface cooling (section 2b). The Gulf of Papua has a big hF i signal corresponding to the season of strongest river outflows, owing to fresh- water fronts clearly visible in Fig. 13. The Solomon and Bismarck Sea results are explained less easily; they have a vertical structure that is strongest just above the base of the mixed layer rather than throughout the surface layer or with surface intensification. In the maps of hF i (Fig. 13), the large values occur near topographic FIG. 12. Single-point PDF of surface frontogenetic tendency 2 SM source regions; the strongest is near the Vitiaz Strait, F (s 5) in the Coral Sea averaged over the months of January (dashed) and June (solid) in the subdomain indicated in Fig. 1. The the coast of New Ireland, and the Woodlark Archipel- associated skewness values are 154 and 175, respectively, that is, ago in the Solomon Sea. These large values are absent highly frontogenetic. away from topography. We infer that the explanation of the unusual vertical structure must come from the to- pographic generation process in some way not yet in the Coral Sea (Fig. 12). For the months of January and known. In the Bismarck Sea, freshwater fronts do con- June, the PDFs have a strong positive skewness (’150). tribute to the F vertical structure because the river The summer and winter maps in Fig. 13 show a winter outflows are maximized in the season between February enhancement in the Coral Sea soup, the coastal areas of and May, but the strongest values seen between May the Bismarck Sea, and the Gulf of Papua. The spatial and August are topography generated. patterns in Figs. 13 and 5 are strikingly similar, and this The role of freshwater forcing in generating SMs is confirms the link between a dynamically active ageo- subtle because in the open ocean it would be repre- strophic secondary circulation (diagnosed by d), front- sented by a positive buoyancy forcing (BS . 0), which if ogenesis, and frontal strength. Similar results are found equated to the effects of surface heating (BS , 0) would in earlier studies (Capet et al. 2008b), although lacking result in shallower mixed layers, reducing the APE the clear statistical picture provided here. reservoir [(5)] and thereby suppressing SM mixed layer

25 22 FIG. 13. (top) Maps of surface F (s ) and j=hbj (s ), temporally averaged over the (left) summer (JFM) and (right) winter (JJA) seasons over the southwestern Pacific region. (bottom) Seasonal cycle of surface F and j=hbj averaged over the subdomains defined in Fig. 1. These are from the case with Dx 5 1.5 km.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1233

xy 25 FIG. 14. Seasonal cycle of the frontogenetic tendency hF i (s ), with spatial averaging performed over the indicated subdomains, from the Dx 5 1.5 km run. Locations with depths shallower than 200 m are excluded. For comparison, the associated surface buoyancy flux hBoi is also plotted (blue dashed line) along with Hb (black line).

generation. However, localized river outflows add fresh shallower Hb because of stronger restratification fluxes (and sometimes cool, like the Mississippi River) water that are absent in the former case. Similar results are into a saltier (and warmer) ocean creating large salinity found in the present study with generally shallower Hb and density fronts. Figure 13 highlights riverine fronto- values found in the Dx 5 1.5 km case relative to the Dx 5 F xy genesis in the form of enhanced values of h im and 4 km case (Fig. 16). During the winter months, Hb in the = xy hj hbj im in the Gulf of Papua, and to a lesser extent in Solomon Sea (the green curve) is shallower, and the the Bismarck region, from January to April when the shallowing in the Dx 5 500 m case is particularly dra- riverine output is largest. In fact, the freshwater frontal matic compared to the lower-resolution cases by a factor gradients are as strong as those generated in the winter of about 2. The Couvelard et al. (2015) study, however, SM soup in the Coral Sea region. Our results mirror the lacks wind stress and a corresponding parameterized findings of Luo et al. (2016) for the Mississippi River turbulent mixing that can have also have a significant even though the total annual output there is roughly SM imprint (Nagai et al. 2006). Idealized studies show twice the Gulf of Papua river outflows. Figure 15 illus- that mixing can enhance MLE-induced restratification trates the impact of freshwater outflows at the ocean surface of this region for the month of April with strong values of hj=hbjim and hF im. Incidentally, while the Gulf of Papua rivers flow into the shallow shelf regions (like the Mississippi), the rivers to the north do not (Fig. 1).

4. Mixed layer submesoscale processes a. Mixed layer depth and APE

The mixed layer depth Hb is diagnosed using the standard approach of the depth at which the tempera- ture is 0.28C below that at the surface. The depth Hb is principally determined by three physical processes: 1) the surface buoyancy flux that causes convective mixing-induced deepening or shallowing depending on the sign of the flux, 2) turbulent boundary layer mixing (that would in general also include surface wave effects) parameterized here by KPP, and 3) restratification fluxes induced by SM processes. Couvelard et al. (2015) attempts to quantify the impact of the SM fronts and eddies on Hb in an idealized baroclinic jet subjected to a winter cooling period. By varying their model resolution 25 FIG. 15. Maps of S (psu) and frontogenetic tendency F (s ) from an eddy-resolving Dx 5 10 km to a SM-permitting averaged over the month of April (the month of peak river outflow Dx 5 2 km, they find that the latter case had significantly in the Gulf), from the Dx 5 1.5 km run.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1234 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

FIG. 16. (top) Seasonal cycle of the mixed layer depth Hb (m) in the Dx 5 4 (black) and 1.5-km (blue) cases, along 2 23 with Bo (m s ) (magenta dashed) for comparison. The Dx 5 500 m Solomon Sea case is also plotted (green) for the 2 winter months. (bottom) APE from (5) (m2 s 2; black) and the mesoscale magnitude of the buoyancy gradient 22 j=bMEj (s ; dashed red).

(Bachman and Taylor 2016) and allow the formation of SMs in earlier sections. Interestingly, the j=hbMEj signal TTW fronts and filaments, which also restratify the here also has a broad winter maximum at odds with the surface layer (McWilliams et al. 2015; McWilliams EKE summer maximum detailed in Qiu et al. (2009) and

2016). Because the seasonal cycles of t and Bo are so briefly discussed in section 3a.Thisisnotinconsistent tightly coupled in the tropics, it can be difficult to de- with geostrophy because mesoscale eddies have a couple their relative effects on Hb variability and asso- thermocline-scale thermal wind balance and not an ciated SMs (more on this in section 3d). Contrary to the Hb-scale balance. The Gulf of Papua region has a prom- discussion above, in previous studies Hb has been found inent February–April maximum in j=hbMEj,owingto to be the strongest determinant of the strength of re- strong river outflows. However, the seasonal cycle stratification fluxes and their seasonal variability, both in of Hb in the Gulf of Papua is out of phase with that of idealized models (Brannigan et al. 2015; Couvelard et al. j=hbMEj, ensuring a broad winter maximum in APE } 2 2015) and the Gulf Stream region (Mensa et al. 2013). because APE Hb is the stronger dependence. Thus, in These results can be understood in the context of the spite of the large surface frontogenesis signal found in mesoscale eddy APE reservoir available to energize SM Fig. 13, shallower mixed layers, caused by strong sur- currents, estimated under assumption of a vertically face heating (Fig. 2), reduce the total APE associated uniform surface buoyancy layer (Fox-Kemper et al. with the riverine output. This has obvious implications 2008; Brannigan et al. 2015)as for the restratification flux. b. Vertical buoyancy fluxes and restratification ’ 1 2 = z APE Hbj b j , (5) 2 h ME The vertical eddy buoyancy flux (VBF), averaged over the mixed layer, represents a transfer of APE to KE where bME is the mesoscale component calculated with a near the surface: spatial low-pass filter whose filter scale is close to R , and d ð the depth average is computed over the mixed layer. 1 xy APE / KE 5 VBF(z) dz, VBF 5 hw0b0 i . Note that Rd increases rapidly close to the equator and a m Hb single filter choice for all latitudes risks corrupting the (6) mesoscale signal with the SM =hb values, which can have large local values due to frontogenesis. We pick the filter Turbulent convection induced by surface cooling would scale to be Lf 5 90 km, somewhat less than the Rd value have negative values (represented as a subgrid-scale of 100 km in the Coral Sea, keeping in mind that lower- process in the ROMS model by KPP), while restratifying latitude values could include some of the SM =hb. SM dynamics that arise from baroclinic conversion The seasonal variability of APE is plotted in Fig. 16 (section 1) have positive values (Boccaletti et al. 2007; for the four regions. The Coral Sea APE has a strong Fox-Kemper et al. 2008; Capet et al. 2008b). Figure 17 winter maximum in line with the statistical measures of shows the seasonal cycle of vertical structure of the

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1235

2 23 FIG. 17. Seasonal cycle of the vertical buoyancy fluxes VBF(z)(m s ), with horizontal averaging over the indicated subdomains. For comparison, the associated surface flux hBoi is plotted (purple) along with Hb (black). spatially averaged VBF in the four subdomains. In study, suggesting a similar dynamical mechanism in the general, VBF is restratifying and leads to a shallowing of two cases. H ; notice in Fig. 16 the shallower values with higher b c. T and S contributions to the VBF resolution, leading to a bigger SM VBF. The Coral Sea displays a clear winter maxima corre- Seawater density depends nonlinearly on temperature sponding to the seasonal cycles of Bo and t. The T, salinity S, and pressure p. However, in the tropical Bismarck and Solomon Seas have weaker seasonal cy- surface layer it is sufficient to approximate it by a linear cles, with the Bismarck Sea in particular showing sum- equation of state mer and spring restratification that coincides with the 2 5 a 2 2 b 2 freshwater forcing from the river outflows and the usual b b0 g(T T0) g(S S0), (8) (but weaker) surface cooling in the winter months. Further, as discussed in section 3b(2), the SM eddies in where b0, S0, and T0 are regional reference values, and a b the Bismarck Sea are mainly coastally and topographi- and are regional constants for the expansion co- cally generated. Gula et al. (2015) analyzed SM gener- efficients. The VBF can therefore be decomposed as ation in the wake of Gulf Stream separation off the w0b0 5 gaw0T0 2 gbw0S0 , (9) Florida Straits and found that the dominant SM energy |fflfflfflffl{zfflfflfflffl} |fflfflffl{zfflfflffl} conversion was from the mean to eddy kinetic energy VTF 2VSF through the horizontal Reynolds stress: where the averaging operator can include both tem- 5 0y 0 1 HRS u uy similar terms, (7) poral and spatial averages. Figures 18 and 19 display the individual contributions from T [i.e., vertical tem- and not VBF. Our results mirror those of Gula et al. perature flux (VTF) 5 gaw0T0]andS [i.e., vertical sa- (2015), but we defer this discussion to section 5. linity flux (VSF) 5 2gbw0S0]totheVBF.Inmost The Gulf of Papua has a VBF that is at its maximum regions, the seasonal patterns of VSF and VBF are during the winter months of May–July (MJJ) and not significantly different. Some of these differences can February–April (FMA), which corresponds to the sea- be related to local forcing effects. The Bismarck and son of strong surface frontogenesis (Fig. 13). The APE Solomon regions, for example, have an expected winter seasonal cycle for this region, estimated in the previous maximum in VTF and a FMA maximum in VSF, re- section, is therefore a better predictor of the VBF than flecting the season of maximum riverine outflow. The xy the surface hF i seasonal cycle, supporting the de- Coral Sea also has a June–August (JJA) maximum that pendence of VBF on Hb. Further, fronts in shallower accounts for the VBF signal during this season. How- mixed layers have weaker, associated vertical velocity, ever, the VSF is largest during June and smaller at again resulting in reduced VBF. Strong summer riverine other times, despite lacking localized freshwater sour- frontogenesis and weak accompanying VBF was found ces during this month (BS ’ 0). To understand this in the Mississippi River outflow region of the Gulf of result, consider the time-averaged equilibrium scalar Mexico by Luo et al. (2016) using a ROMS-based model balance equations for T and S:

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1236 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

xy 5 a 0 0 2 23 FIG. 18. Seasonal cycle of the vertical temperature flux VTF(z) hg w T im (m s ), averaged over the indicated subdomains. For comparison, the associated surface flux hBTi is plotted (purple) along with Hb (black).

= 1 › 5 › k› huhfT, Sgim zhwfT, Sgim zh zfT, Sgim, (10) with KPP. This analysis is consistent with the co- incidence of the seasonal maxima in VSF and t in the where uh 5 (u, y) is the horizontal velocity, and k is the Coral Sea. We hypothesize that this is an example of KPP-supplied diffusivity. A similar equation is true for SM generation by vertical boundary layer mixing in- buoyancy because of the linear approximation in (8). ducedbywindstressthroughTTWfrontsandfila- The horizontal scalar fluxes are dominated by mesoscale ments (McWilliams et al. 2015)andMLEsinaTTW eddies while the vertical fluxes are dominated by SM balance (Bachman and Taylor 2016). Actual justifi- processes. The dominant SM balances are between the cation of this hypothesis would require a careful vertical advection term on the left side and the mixing analysis of the dynamical balance in the surface layer term on the right side (Capet et al. 2008a). because the lowest-order balance of TTW fronts in- On spatially averaging over the subdomains in Fig. 1 volves the vertical mixing term. For now, we merely and integrating between the surface (z 5 0) and mixed note that the restratification flux due to TTW fronts layer depth (z 52Hb), we get [see (20) and the discussion surrounding it] has the

xy form 0 0 ’ k› xy 2 a hw T im h zT im hBT /g im, and (11) xy 0 0 ’ k› xy 1 b 0 0 } hw S im h zS im hBS/g im . (12) w bTTW u*, (13)

Thus, when BS ’ 0, as in the Coral Sea during the and a local maximum in wind stress could lead to a local xy xy 0 0 ’ k› winter months, then hw S im h zS im,andconse- maximum in VBF (through VSF), consistent with the quently VSF } u*duetok } u*inawind-drivenregime results in Figs. 19 and 17.

xy 5 2 b 0 0 2 23 FIG. 19. Seasonal cycle of the vertical salinity flux VSF(z) h g w S im (m s ), averaged over the indicated subdomains. For comparison, the associated surface flux hBSi is plotted (purple) along with Hb (black). Notice that the color bar and BS axis ranges are exactly half of the corresponding values in Figs. 17 and 18.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1237

0 0 0 0 0 0 2 23 FIG. 20. Seasonal cycle of VBF, compared with the scale estimates from MLE w b MLE, TTW w b TTW, and a frontogenesis w ba (m s ) integrated over Hb for the Dx 5 1.5 km case only, with spatial averages performed over the indicated areas. d. Restratification flux parameterizations aH2j= b j F;2 b h ME 2 . (16) Fox-Kemper et al. (2008) proposed a parameterization f for restratification due to MLEs, under the assumption The corresponding restratification flux averaged over of a constant N2 and = b over the mixed layer depth h ME many such fronts is [similar to APE expression in (5)]. This is expressed in terms of an ageostrophic overturning eddy-induced aH2j= b j2 0 0 b h ME streamfunction F*, derived through a combination of w ba ; C , (17) a f 2 model-based phenomenology and heuristics in the form where the best-fit estimate for the southwestern Pacific H2› b F ; b x ME region was Ca 5 2.5, which is consistent with the O(1) * CMLE , (14) f value suggested by McWilliams (2016) in an idealized model of a front and a dense filament. where x is the cross-front direction, and C is an empir- MLE TTW fronts and filaments form under conditions of ical constant. In the mixed layer the VBF can be written in strong vertical mixing produced by a surface wind stress, terms of F*as(Colas et al. 2013; Held and Schneider 1999), in which case the lowest-order balance in the surface H2j= b j layer is no longer geostrophy but w0b0 ;F*j= b j ; C b h ME . (15) MLE h ME MLE f 3 52= f 1 › n › f › f 5 f k uh h z( y z ), z b. (18) Here, we have assumed that the cross-frontal gradient Under such conditions nonzero values of k 3 = b can be dominates j= b j,thatis,› b › b .Equation h h ME x ME y ME enhanced by an emergent ageostrophic secondary cir- (15) has been tested in realistic model configurations culation, which leads to fronts and filaments similar in (Mensa et al. 2013; Luo et al. 2016) with some success but structure to those seen in ‘‘normal’’ frontogenesis. In with caveats about the choice of C . Fox-Kemper et al. MLE this case, the scaling for the overturning streamfunction (2008) diagnoses the value of C in an idealized frontal MLE takes the form (McWilliams 2016, manuscript submitted spindown problem and finds C 5 0.06. Bachman and MLE to J. Fluid Mech.) Taylor (2016) use equilibrated model cases of an ideal- ized horizontal buoyancy gradient in the presence of n = yj b j mixing and suggest O(1) values for the constant. Our F;2 h ME . (19) f 2 value for the best fit shown in Fig. 20 is CMLE 5 0.3, which is closer to the Bachman and Taylor (2016) estimate. Writing the eddy viscosity based on mixing length ar- Mesoscale eddies can themselves generate fronts in guments or KPP as ny 5 u*Hb, we can write the re- the surface mixed layer, although their strain-induced stratification flux in the form of frontogenesis occurs when a mesoscale strain field = 2 enhances a favorably aligned hbME through an ageo- u*H j= b j 5 2F F w0b0 ; C b h ME , (20) strophic secondary circulation (u, w) ( z, x), with TTW TTW f 2 F as the associated overturning streamfunction. For a 2 2 given mesoscale strain field a, where a 5 (ux 1 yy) 1 where the best-fit estimate for CTTW in this region is 2 (uy 2 yx) , the overturning streamfunction for a front CTTW 5 0.07, which is close to the ’0.1 value found in the takes the form nonlinear TTW-front solutions by McWilliams (2016).

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1238 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

There is a striking similarity in the forms of the three where restratification expressions (15), (17), and (20). In fact, [2 0 0 1 0y0 1 y0y0y 1 0y0y all three dynamical processes extract energy from the HRS (u u ux u uy y u x) (22) common APE reservoir derived in (5). Depending on the relative strengths of a, u*, and the depth of mixed and layer Hb, one or another process might dominate. Thus, 52 y0 0y 1 0 0 during the summer when the ML is shallow and under VRS ( w z u w uz) (23) weak wind forcing conditions, w0b0a might be relevant, while the other two might coexist in conditions of are the contributions from the horizontal and vertical strong winds and surface cooling, like in the south- Reynolds stresses to the mean to eddy transfer; KmKe . western Pacific. Based on the discussion in section 4c, 0 implies that energy is transferred from the mean to the we might expect wind stress effects to be more impor- eddies and vice versa. In general, topographic eddy tant in the Coral Sea during June and cooling after. generation can occur at both mesoscales and SMs with However, based purely on the scalings derived above, the eddy scales depending on the Rossby and Burger the attribution problem is a difficult one, owing to the [Bu 5 (NH/fL)2, where N is the local buoyancy fre- similar manifestation of frontal features in the three quency] numbers associated with the topographic - cases. In Fig. 20, we match VBF values obtained from stacles themselves (Dong and McWilliams 2007). ROMS with the three scalings. The constants (one for Figure 21 shows a spatial map of HRS at a depth of each scaling) are chosen to match the three geo- z 5220 m (which is well within the mixed layer in this graphical regions. While all three scalings do reason- region) over the month of June [thus the averaging op- ably in the Solomon and Coral Seas, there is a erator (22) is a monthly average here]. Strong positive pronounced lag of about 2 months in the Gulf of Papua signals in the wakes of headlands and island wakes sur- subdomain, which is likely related to the interplay of rounding the Solomon and Bismarck Seas are seen. shallow mixed layers and strong surface fronts that was These patches of positive HRS (mean to eddy conver- speculated on in sections 4a and 4b. Equation (20) implies sions) have narrow (cross flow) spatial structures that all three parameterizations can be reasonable fits for of ,20 km in width, indicating that the eddy generation w0b0, which relates to the similarity in their scalings and due to barotropic conversions happens at the SMs. their primary energy source in surface layer APE. Con- These are likely a consequence of the small island sizes sequently, attempts to attribute all restratification to and equatorial proximity as conjectured in section 3b. MLEs (Mensa et al. 2013; Luo et al. 2016) are doubtful. The HRS signals in the inland Solomon Sea and the While instructive, the derived scalings are insufficiently Coral Sea (not shown) are considerably weaker, seen in discriminating among the mixed layer SM processes, and the figure as faint ‘‘smudges’’ of mesoscale extent. Based thus other approaches must be used. on the visual patterns of HRS distribution in this region, we examine the seasonal cycle of the energy conversions in four topographic regions marked as (1)–(4) in Fig. 21. 5. Topographic submesoscale wakes Following the analysis methodology in preceding sec- It is indicated in section 3b(2) that the interaction of tions, we spatially average the three-dimensional HRS, surface waters with the various island chains in and VRS, and VBF over the respective regions to view their around the Bismarck and Solomon Seas is a significant seasonal cycles over the top 100 m. Figure 22 has the source of SCVs. These SCVs are formed by a results and can be used to infer the nature of the SM barotropic–centrifugal instability of vortical wakes generation in these regions. In region 1, along the north through separation of bottom drag–generated shear coast of Papua New Guinea, HRS ; 3VBF ; 10VRS in layers (Molemaker et al. 2015). To characterize the the winter months, which shows that barotropic in- flow–topographic interactions, we examine the ener- stability is dominant, as seen in the HRS patterns in getics of the surface flow. Barotropic instabilities Fig. 21 primarily along topographic wakes. The winter involve a transfer of energy from the background mean peak in HRS is contemporaneous with the peak mass flow to the eddies. With a decomposition of the flow flow rate across the Vitiaz Strait (not shown). Further, fields into mean and deviations as u 5 u 1 u0, where the we attribute the VBF signal as being due to mixed layer bar represents a combination of spatial and temporal SM generation rather than mixed barotropic–baroclinic averages, the energy transfer from mean to eddy KE is instability, mediated by topography. Regions 2 and 3, written as (Dong and McWilliams 2007) composed of the island chains on the eastern coast of the Solomon Sea, have HRS ; 2VBF ; 10VRS, whereas in KmKe 5 HRS 1 VRS, (21) the southwestern Solomon Sea, VBF is of greater

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1239

2 23 FIG. 21. HRS (m s )atz 5220 m averaged over the month of June for the topographically active region in and around the Solomon and Bismarck Seas. The four marked regions are the subject of further analysis in Fig. 22. importance. Region 4, where the western boundary Papua New Guinea coast, both relative to mixed layer current traverses the Louisiade Archipelago at depth, SM processes and to other topographically complex does not have a strong SM generation signal, with only regions surrounding the Solomon Sea. VRS a modest source in the surface layer. Thus, the It should be noted that the choice of the monthly av- strongest topographic generation happens along the eraging operator only identifies energy transfers from

2 23 FIG. 22. Seasonal cycles of HRS, VRS, and VBF (m s ) spatially averaged over the four regions marked in Fig. 21. Each row represents a region, with region 1 in Fig. 21, the first row, region 2 in the second row, and so on. Notice that the color bar ranges vary both among the rows and columns. The black curve in each plot shows the variation of Hb, spatially averaged over the respective region.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1240 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47 mean flows and mesoscale eddies with lifetimes longer Atlantic and finds a close association with the formation, than a month, and the transfer of energy from shorter- and in some cases the enhancement, of freshwater lived mesoscale eddies to SCVs would be unac- fronts. They find that the barrier layers have a small counted for. enhancement effect on the salinity restratification flux (VSF). Here, we have not examined the barrier layer dynamics in detail, partly because, unlike the North 6. Summary and conclusions Atlantic, the topographic SMs in our region of interest Surface layer submesoscale currents (SMs) are ener- dominate other effects and also because our climato- gized by a combination of nonlocal mesoscale eddy ef- logical E 2 P forcing might underestimate the effects of fects (through =MEb and a) and local forcings (t and Bo) episodic high rain events. Flow–topographic in- that generate surface layer APE [(5)] and vertical mix- teractions convert KE of the mean flow to KE in me- ing. The APE reservoir can be accessed by any of the soscales and SMs, depending on the Rossby numbers three mechanisms of TTW, MLEs, or strain-induced associated with the topographic obstacles (Dong and frontogenesis that are manifested through VBF, the McWilliams 2007). For large values of Ro, as is the case conversion of mesoscale APE to frontal-scale KE, of the headlands and island chains around the near- generating SMs and concomitantly restratifying the equatorial (f / 0) eastern and western coasts of mixed layer. In the Coral Sea the restratification flux the Bismarck Sea, the generation is largely in the SM has a broad peak during the winter months of JJA. Be- range. These topography-induced SMs have weak re- cause no significantly localized river sources of fresh- stratification flux VBF, as also is seen in idealized (Dong water exist, a decomposition of the restratification flux and McWilliams 2007) and realistic midlatitude models into T and S components shows distinct peaks related to (Gula et al. 2014), but they do have a strong signature in the months of maximum wind stress (seen in the vertical SM statistical measures of vertical velocity and vorticity salinity flux) and surface cooling (seen in the vertical variance. Moreover, the conversion of kinetic energy temperature flux). These two peaks are also seen in the from the mean to eddies by horizontal Reynolds stress seasonal cycle of frontal tendency. Equatorward of the HRS is larger than that due to the vertical Reynolds Coral Sea, both the seasonal cycle and the peak value of stress VRS (which is negligible at the SM) or the vertical

Bo are weak, as are the corresponding restratification buoyancy flux VBF, indicating a dominant barotropic effects. The eastern Solomon Sea, in particular, is pop- instability mechanism. North of the Vitiaz Strait, along ulated by large, Rd-size mesoscale eddies that are not the coast of Papua New Guinea, the HRS terms are conducive to SM frontogenesis, leading to a relative SM about 3 times larger than the VBF, while along the desert in this region year-round, clearly observed in the Solomon Islands, they are larger by about a factor of 2. SM statistics of vorticity, surface divergence, and Furthermore, the HRS conversion primarily generates buoyancy gradient. Freshwater forcing at river mouths SCVs in these regions. along the northern coast of Papua New Guinea and the A key observation in this paper is that in spite of western Solomon Sea generates fronts that contribute to their near-coastal localization, topography-induced sub- the restratification flux and are comparable in magnitude mesoscales can be, on average, stronger than those in- to the VBF during the winter deepening. Restratification duced through mixed layer instabilities, evident from induced by freshwater fronts and mixed layer fronts, the zonally averaged statistics presented in Fig. 11.The following the winter deepening, display distinct signa- dynamical consequences of this result remain to be tures in VSF and VTF, respectively. The largest riverine explored. A possible effect is transfer of energy to outflow in the region is in the Gulf of Papua, resulting in smaller scales through three-dimensional centrifugal strong surface frontogenesis that matches the Coral Sea instability (CI), a form of negative potential vorticity SM soup in intensity. However, there is only a weak re- instability. Molemaker et al. (2015) showed enhanced stratification effect due to these freshwater fronts in the local dissipation in the California Undercurrent as it Gulf of Papua, possibly due to shallow mixed layers in the separates off a headland and generates anticyclonic summer season when the river outflows peak. Shallow SCVs, suggesting a CI-induced forward cascade (Dewar mixed layers reduce the APE reservoir and subsequently et al. 2015; Jiao and Dewar 2015). This is likely to be of the restratification fluxes. particular importance in the Bismarck and Solomon An important aspect of tropical that has not Seas, where anticyclones with large vortex Rossby been discussed here is the role of barrier layers on the number Ro ; 10 are commonplace [see section 3b(2)]. SMs, that is, haloclines that limit the depth of surface The Dx 5 500 m and subsequently planned higher- layer vertical mixing. A recent study by Veneziani et al. resolution runs are likely to shed greater light on (2014) focuses on boundary layers in the tropical North these phenomena.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC JUNE 2017 S R I N I V A S A N E T A L . 1241

Acknowledgments. This work was funded by NASA Fox-Kemper, B., R. Ferrari, and R. Hallberg, 2008: Parameterization Grant NNX13AP51G and NSF Grant OCE-1419450. of mixed layer eddies. Part I: Theory and diagnosis. J. Phys. Oceanogr., 38, 1145–1165, doi:10.1175/2007JPO3792.1. Gourdeau, L., J. Verron, A. Melet, W. Kessler, F. Marin, and REFERENCES B. Djath, 2014: Exploring the mesoscale activity in the Solomon Sea: A complementary approach with a numerical model and Bachman, S. D., and J. R. Taylor, 2016: Numerical simulations of altimetric data. J. Geophys. Res. Oceans, 119, 2290–2311, the equilibrium between eddy-induced restratification and doi:10.1002/2013JC009614. vertical mixing. J. Phys. Oceanogr., 46, 919–935, doi:10.1175/ Gula, J., M. J. Molemaker, and J. C. McWilliams, 2014: Sub- JPO-D-15-0110.1. mesoscale cold filaments in the Gulf Stream. J. Phys. Ocean- Boccaletti, G., R. Ferrari, and B. Fox-Kemper, 2007: Mixed layer ogr., 44, 2617–2643, doi:10.1175/JPO-D-14-0029.1. instabilities and restratification. J. Phys. Oceanogr., 37, 2228– ——, ——, and ——, 2015: Topographic vorticity generation, 2250, doi:10.1175/JPO3101.1. submesoscale instability and vortex street formation in the Brannigan, L., D. P. Marshall, A. Naveira-Garabato, and A. G. Gulf Stream. Geophys. Res. Lett., 42, 4054–4062, doi:10.1002/ Nurser, 2015: The seasonal cycle of submesoscale flows. Ocean 2015GL063731. Modell., 92, 69–84, doi:10.1016/j.ocemod.2015.05.002. Held, I. M., and T. Schneider, 1999: The surface branch of the zonally Callies, J., R. Ferrari, J. M. Klymak, and J. Gula, 2015: Seasonality averaged mass transport circulation in the troposphere. J. Atmos. in submesoscale turbulence. Nat. Commun., 6, 6862, Sci., 56, 1688–1697, doi:10.1175/1520-0469(1999)056,1688: doi:10.1038/ncomms7862. TSBOTZ.2.0.CO;2. Capet, X., J. C. McWilliams, M. J. Molemaker, and A. F. Hoskins, B. J., and F. P. Bretherton, 1972: Atmospheric fronto- Shchepetkin, 2008a: Mesoscale to submesoscale transition in genesis models: Mathematical formulation and solution. the California Current System. Part I: Flow structure, eddy J. Atmos. Sci., 29, 11–37, doi:10.1175/1520-0469(1972)029,0011: flux, and observational tests. J. Phys. Oceanogr., 38, 29–43, AFMMFA.2.0.CO;2. doi:10.1175/2007JPO3671.1. Hristova, H. G., W. S. Kessler, J. C. McWilliams, and M. J. ——, ——, ——, and ——, 2008b: Mesoscale to submesoscale tran- Molemaker, 2014: Mesoscale variability and its seasonality in sition in the California Current System. Part II: Frontal processes. the Solomon and Coral Seas. J. Geophys. Res. Oceans, 119, J. Phys. Oceanogr., 38, 44–64, doi:10.1175/2007JPO3672.1. 4669–4687, doi:10.1002/2013JC009741. Carton, J. A., and B. S. Giese, 2008: A reanalysis of ocean climate Jiao, Y., and W. Dewar, 2015: The energetics of centrifugal in- using Simple Ocean Data Assimilation (SODA). Mon. Wea. stability. J. Phys. Oceanogr., 45, 1554–1573, doi:10.1175/ Rev., 136, 2999–3017, doi:10.1175/2007MWR1978.1. JPO-D-14-0064.1. Chavanne, C. P., and P. Klein, 2016: Quasigeostrophic diagnosis of Kessler, W. S., and S. Cravatte, 2013: ENSO and short-term vari- mixed layer dynamics embedded in a mesoscale turbulent ability of the South Equatorial Current entering the Coral Sea. field. J. Phys. Oceanogr., 46, 275–287, doi:10.1175/ J. Phys. Oceanogr., 43, 956–969, doi:10.1175/JPO-D-12-0113.1. JPO-D-14-0178.1. Large, W., J. C. McWilliams, and S. C. Doney, 1994: Oceanic Chelton, D. B., M. G. Schlax, and R. M. Samelson, 2011: Global vertical mixing: A review and a model with a nonlocal observations of nonlinear mesoscale eddies. Prog. Oceanogr., boundary layer parameterization. Rev. Geophys., 32, 363–403, 91, 167–216, doi:10.1016/j.pocean.2011.01.002. doi:10.1029/94RG01872. Colas, F., X. Capet, J. C. McWilliams, and Z. Li, 2013: Mesoscale Lemarié, F., J. Kurian, A. F. Shchepetkin, M. J. Molemaker, eddy buoyancy flux and eddy-induced circulation in eastern F. Colas, and J. C. McWilliams, 2012: Are there inescapable boundary currents. J. Phys. Oceanogr., 43, 1073–1095, issues prohibiting the use of terrain-following coordinates in doi:10.1175/JPO-D-11-0241.1. climate models? Ocean Modell., 42, 57–79, doi:10.1016/ Couvelard, X., P. Marchesiello, L. Gourdeau, and J. Lefèvre, 2008: j.ocemod.2011.11.007. Barotropic zonal jets induced by islands in the southwest Pacific. Luo, H., A. Bracco, Y. Cardona, and J. C. McWilliams, 2016: J. Phys. Oceanogr., 38, 2185–2204, doi:10.1175/2008JPO3903.1. Submesoscale circulation in the northern Gulf of Mexico: ——, F. Dumas, V. Garnier, A. Ponte, C. Talandier, and Surface processes and the impact of the freshwater river input. A. Treguier, 2015: Mixed layer formation and restratification Ocean Modell., 101, 68–82, doi:10.1016/j.ocemod.2016.03.003. in presence of mesoscale and submesoscale turbulence. Ocean Mason, E., M. J. Molemaker, A. Shchepetkin, F. Colas, J. C. Modell., 96, 243–253, doi:10.1016/j.ocemod.2015.10.004. McWilliams, and P. Sangrà, 2010: Procedures for offline grid Danabasoglu, G., W. G. Large, J. J. Tribbia, P. R. Gent, B. P. nesting in regional ocean models. Ocean Modell., 35, 1–15, Briegleb, and J. C. McWilliams, 2006: Diurnal coupling in the doi:10.1016/j.ocemod.2010.05.007. tropical oceans of CCSM3. J. Climate, 19, 2347–2365, McWilliams, J. C., 1985: Submesoscale, coherent vortices in doi:10.1175/JCLI3739.1. the ocean. Rev. Geophys., 23, 165–182, doi:10.1029/ Dewar, W., J. McWilliams, and M. Molemaker, 2015: Centrifugal RG023i002p00165. instability and mixing in the California Undercurrent. J. Phys. ——, 2016: Submesoscale currents in the ocean. Proc. Roy. Soc., Oceanogr., 45, 1224–1241, doi:10.1175/JPO-D-13-0269.1. A472, 20160117, doi:10.1098/rspa.2016.0117. Djath, B., J. Verron, A. Melet, L. Gourdeau, B. Barnier, and J.-M. ——, and E. Huckle, 2006: Ekman layer rectification. J. Phys. Molines, 2014: Multiscale dynamical analysis of a high- Oceanogr., 36, 1646–1659, doi:10.1175/JPO2912.1. resolution numerical model simulation of the Solomon Sea ——, ——, and A. F. Shchepetkin, 2009: Effects in a stratified circulation. J. Geophys. Res. Oceans, 119, 6286–6304, Ekman layer. J. Phys. Oceanogr., 39, 2581–2599, doi:10.1175/ doi:10.1002/2013JC009695. 2009JPO4130.1. Dong, C., and J. C. McWilliams, 2007: A numerical study of island ——,J.Gula,M.J.Molemaker,L.Renault,andA.F.Shchepetkin, wakes in the Southern California Bight. Cont. Shelf Res., 27, 2015: Filament frontogenesis by boundary layer turbulence. 1233–1248, doi:10.1016/j.csr.2007.01.016. J. Phys. Oceanogr., 45, 1988–2005, doi:10.1175/JPO-D-14-0211.1.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC 1242 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 47

Melet, A., L. Gourdeau, and J. Verron, 2010: Variability in Solomon ——, and ——, 2005: The Regional Oceanic Modeling System Sea circulation derived from altimeter sea level data. Ocean (ROMS): A split-explicit, free-surface, topography-following- Dyn., 60, 883–900, doi:10.1007/s10236-010-0302-6. coordinate oceanic model. Ocean Modell., 9, 347–404, Mensa, J. A., Z. Garraffo, A. Griffa, T. M. Özgökmen, A. Haza, doi:10.1016/j.ocemod.2004.08.002. and M. Veneziani, 2013: Seasonality of the submesoscale dy- ——, and ——, 2009: Correction and commentary for ‘‘Ocean namics in the Gulf Stream region. Ocean Dyn., 63, 923–941, forecasting in terrain-following coordinates: Formulation doi:10.1007/s10236-013-0633-1. and skill assessment of the regional ocean modeling sys- Molemaker, M. J., J. C. McWilliams, and W. K. Dewar, 2015: tem’’ by Haidvogel et al., J. Comp. Phys. 277, pp. 3595– Submesoscale instability and generation of mesoscale anticy- 3624. J. Comput. Phys., 228, 8985–9000, doi:10.1016/ clones near a separation of the California Undercurrent. j.jcp.2009.09.002. J. Phys. Oceanogr., 45, 613–629, doi:10.1175/JPO-D-13-0225.1. Stammer, D., 1997: Global characteristics of ocean variability es- Nagai, T., A. Tandon, and D. Rudnick, 2006: Two-dimensional timated from regional TOPEX/POSEIDON altimeter mea- ageostrophic secondary circulation at ocean fronts due to surements. J. Phys. Oceanogr., 27, 1743–1769, doi:10.1175/ vertical mixing and large-scale deformation. J. Geophys. Res., 1520-0485(1997)027,1743:GCOOVE.2.0.CO;2. 111, C09038, doi:10.1029/2005JC002964. Tulloch, R., J. Marshall, C. Hill, and K. S. Smith, 2011: Scales, Qiu, B., S. Chen, and W. S. Kessler, 2009: Source of the 70-day meso- growth rates, and spectral fluxes of baroclinic instability in the scale eddy variability in the Coral Sea and the North Fiji Basin. ocean. J. Phys. Oceanogr., 41, 1057–1076, doi:10.1175/ J. Phys. Oceanogr., 39, 404–420, doi:10.1175/2008JPO3988.1. 2011JPO4404.1. Shchepetkin, A. F., and J. C. McWilliams, 2003: A method for Veneziani, M., A. Griffa, Z. Garraffo, and J. A. Mensa, 2014: computing horizontal pressure-gradient force in an oceanic Barrier layers in the tropical South Atlantic: Mean dynamics model with a nonaligned vertical coordinate. J. Geophys. Res., and submesoscale effects. J. Phys. Oceanogr., 44, 265–288, 108, 3090, doi:10.1029/2001JC001047. doi:10.1175/JPO-D-13-064.1.

Unauthenticated | Downloaded 10/09/21 06:26 AM UTC