1. Introduction
The sea surface density field contains rich variability over submesoscale O(0.1–10) km length scales (e.g., McWilliams 2016) that often manifest as density fronts and filaments. Previous studies have shown that submesoscale fronts and filaments in the open ocean (hundreds of kilometers from shore) can affect the transport of biogeochemical tracers and contaminants (e.g., Franks 1992; Nagai et al. 2015; Mahadevan 2016; Lévy et al. 2018). Submesoscale density fronts are ubiquitous on continental shelves in high-resolution coastal models (e.g., Romero et al. 2016; Dauhajre et al. 2017, 2019), observed within 10 km from shore (e.g., Ohlmann et al. 2017; Connolly and Kirincich 2019), and detected in satellite sea surface temperature (SST) images (e.g., Castelao et al. 2006; Kahru et al. 2012). Dye and SST observations reveal frontal variability within 1 km from shore (Hally-Rosendahl et al. 2015; Grimes et al. 2020). Fronts alter Lagrangian transport pathways and water mass structure over the continental shelf (e.g., Banas et al. 2009; Rao et al. 2011). Here, we focus on the dynamics of submesoscale fronts in the shallow coastal ocean (<10 km from shore and <30-m water depth). In coastal regions at O(0.1–10) km length scales, studies have largely focused on fronts associated with inlet and river plumes (e.g., O’Donnell 2010; Chant 2011; Horner-Devine et al. 2015; Feddersen et al. 2016) and upwelling (e.g., Brink 1987; Austin and Barth 2002; Austin and Lentz 2002). Instead, we focus on coastal fronts in a region of weak winds and no significant freshwater flows, a poorly studied part of the coastal submesoscale frontal parameter space.
A companion paper of this work (Wu et al. 2020, hereinafter W20) investigated the processes transporting shoreline released dye representing wastewater off the San Diego (United States)/Tijuana (Mexico) coast in the San Diego Bight (see Fig. 1) using a high-resolution realistic wave–current coupled model. On the mid- to outer shelf boundary (smoothed 25-m isobath, ≈5 km from shore), wind-driven Ekman transport and submesoscale flows both played an important role in offshore dye transport during a three month analysis period. The submesoscale flows were elevated for stronger root-mean-square (rms) surface alongshore density gradients at length scales < 15 km, which were enhanced by the large scale (over ≈15 km alongshelf) convergent northward alongshore flow, suggesting cross-shore-oriented fronts. In high-resolution numerical models, density gradients are preferentially perpendicular to bathymetric contours in depths < 50 m (Romero et al. 2013; Dauhajre et al. 2017), suggesting alongshore-oriented fronts and filaments. In general, the kinematics (i.e., occurrence likelihood, orientations, lengths, and density gradient magnitude) of coastal (within 10 km of shore) density fronts in regions with little freshwater input are poorly understood.

LV4 grid bathymetry (color shading) and the front study region (white line) to which mean front locations are restricted. Red dots denote the freshwater sources Punta Bandera (PB), Tijuana River Estuary (TJRE), and Sweetwater River. The yellow dot denotes the South Bay Ocean Outfall (SB) mooring site in 30-m depth. San Diego Bay (SDB), Point Loma, and the U.S.–Mexico border are also labeled.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
The DF mechanism involves a large-scale geostrophic, nondivergent strain field whose cross-front convergence enhances the density gradient and accelerates an alongfront jet, which via Coriolis forcing induces an ageostrophic cross-front flow υa (Hoskins and Bretherton 1972; Hoskins et al. 1978). The induced υa and the associated downwelling and upwelling form an ageostrophic secondary circulation (ASC) tilting the isopycnals toward the horizontal (e.g., Bleck et al. 1988; Spall 1995; Thomas et al. 2008). TTW refers to a balance among vertical mixing, Coriolis and pressure gradient forcing (McWilliams et al. 2015), where the ageostrophic Coriolis forcing is balanced by vertical mixing (Garrett and Loder 1981; Thompson 2000; Gula et al. 2014; Wenegrat and McPhaden 2016). Cross-front varying vertical mixing of momentum [i.e., ∂z(Aυ∂zu), where Aυ is the vertical eddy viscosity and u is the alongfront velocity] can induce a cross-front ageostrophic convergence ∂υa/∂y ≠ 0, enhancing the density gradient and forming a TTW ASC (McWilliams 2017). In numerical models in the Gulf of Mexico, the density gradient of TTW generated fronts and filaments can strengthen rapidly on hourly time scales consistent with an asymptotic model assuming weak near-surface stratification (Barkan et al. 2019). The TTW mechanism was invoked to explain the strengthening of coastal density filaments and fronts during winter and spring in a high-resolution numerical model (Dauhajre et al. 2017). However, to what extent the DF and TTW mechanisms are generally applicable to generation of density fronts in coastal regions within 10 km from shore is unclear. Stratification and wind forcing in coastal regions vary dramatically. The spatial variability of horizontal density gradient varies seasonally in the California Current System (CCS; Kahru et al. 2012; Mauzole et al. 2020) and in the Gulf Stream (Callies et al. 2015). The shoreline limits the onshore extent of fronts, the shore normal velocity vanishes at the shoreline, and shallow coastal depths constrain frontal vertical circulation. The shoreline also constrains shelf circulation to largely a geostrophic (ageostrophic) balance in the cross-shore (alongshore) direction (e.g., Allen 1980; Lentz et al. 1999).
Here, we focus on the kinematics and dynamics of coastal density fronts (within 10 km from shore and <30-m depth) using a high-resolution numerical model of the San Diego Bight (W20). A field study in this region noted the enhancement of a dye alongshore front driven by the internal tide (Grimes et al. 2020). In this region, winds are relatively weak and freshwater input is small, placing focus on unforced (e.g., DF and TTW) frontogenesis mechanisms. An edge detection method is used to isolate individual density fronts. We address three main questions. What are the kinematic properties (i.e., orientation, length and density gradient) of these coastal fronts? For cross-shore-oriented fronts, what does a typical front look like? What are the processes responsible for the frontogenesis and can they be classified in the context of open ocean unforced frontogenesis mechanisms? The model configuration, front detection procedure, and front kinematic parameters are given in section 2. Front kinematic properties and variability are analyzed in section 3. An ensemble mean cross-shore-oriented front is created to quantify frontal circulation in section 4. Frontogenesis mechanisms are diagnosed through frontogenesis tendency and a momentum balance analysis in sections 5 and 6, respectively. Front dynamics in the context of DF and TTW mechanisms are discussed in section 7. A summary is provided in section 8.
2. Model configuration and front detection
a. Model setup
Shelf and surfzone circulation is simulated using the Coupled Ocean–Atmosphere–Wave–Sediment–Transport (COAWST) model system (Warner et al. 2010; Kumar et al. 2012). A full description of the model setup is found in W20. Here only the information essential to this work is provided. The model consists of three one-way nested parent runs (from LV1 to LV2 and then LV3) spanning from the California Current System to the south Southern California Bight, and one downscaled high-resolution child run (LV4) resolving the outer to inner shelf and surfzone in the southern San Diego Bight (Fig. 1). LV4 incorporates surface waves by coupling the Regional Ocean Modeling System (ROMS; Shchepetkin and McWilliams 2005) with the Simulating Waves Nearshore model (SWAN; Booij et al. 1999). NOAA/NAM surface fluxes (wind stress, heat, and precipitation) are applied. Vertical mixing (eddy viscosity and diffusivity) is derived from a k–ϵ submodel (e.g., Umlauf and Burchard 2003) with Kantha and Clayson (1994) stability functions. In all simulations, a third-order upwind advection scheme is used for momentum. The horizontal eddy viscosity and diffusivity are constant at 0.5 m2 s−1 over all the model runs. For the LV4 grid, this horizontal eddy viscosity and diffusivity have little effect on submesoscale variability (W20).
The LV4 grid (15 × 36 km2) spans from Punta Bandera (PB), Mexico, to Point Loma, United States, encompassing the Tijuana River Estuary (TJRE) and the San Diego Bay (SDB) (Fig. 1). The shoreline is relatively straight, except for curvature around SDB and a broad 15-m depth shoal offshore of the TJRE mouth. The horizontal grid resolution transitions from 100 m along the three open boundaries to 8 m approaching the TJRE mouth, resulting in a regional mean resolution ≈30 m. The vertical stretched grid has 15 s levels with enhanced resolution near the surface and bottom. The number of vertical levels is limited to prevent thin vertical layers in very shallow (<1 m) depths. As we are focused on surface density fronts, we provide context of vertical grid resolution. In 30-m depth, the average vertical resolution is Δz = 0.8 m for z > −5 m and for −10 < z < −5, the average vertical resolution is Δz = 2 m. The initial and boundary conditions, nested from the parent LV3 solution, include both barotropic and baroclinic tides. Barotropic tides are prescribed on the outmost LV1 grid, allowing for the generation of baroclinic tides within all model domains (e.g., Kumar et al. 2015; Suanda et al. 2017; Kumar et al. 2019). LV4 receives realistic freshwater discharge from PB, TJRE, and the Sweetwater River within SDB. TJRE discharge occurs following intermittent rainfall events. At PB, untreated wastewater outflows are represented with a constant freshwater discharge (Qr = 1.53 m3 s−1; see W20 for more details). The simulation is conducted from July to October 2015 using XSEDE resources (Towns et al. 2014), and solutions are saved at 1-h intervals.
b. Regional oceanographic conditions
Following W20, model results are analyzed over the summer to fall transition (22 July–18 October 2015, denoted analysis period). The barotropic mixed tides have an amplitude around 1 m (Fig. 2a). NAM winds are mostly southeastward directed and have a low (|Uw| < 5 m s−1) to moderate (5–8 m s−1) speed (Fig. 2b). The shelf stratification is represented by the top-to-bottom buoyancy frequency N2 = −(g/ρ0)Δρ/Δz at a central location denoted SB (30-m depth, see Fig. 1 for location), where g is gravity and the background density ρ0 = 1025 kg m−3. The subtidal (low-pass filtered with a 33-h cutoff) N2 decreases overall from a relatively strong 5 × 10−4 s−2 during summer to 1 × 10−4 s−2 during fall (Fig. 2c), typical for summer to fall stratification in this region of Southern California (e.g., Palacios et al. 2004). Within the LV4 grid, the time-mean surface density has a weak north–south gradient reaching a midshelf (25-m isobath) magnitude of 6 × 10−6 kg m−4 with lighter water to the north (W20). This is due to the regional differences in upwelling between Southern California and Baja California (e.g., Huyer 1983), which is also seen in the parent LV3 grid (W20). The SDB has negligible freshwater input during this time period (W20); however, the warm water of the SDB serves as a weak buoyancy source, which may slightly augment this already present north–south regional density gradient. The subtidal depth-averaged alongshore flow at SB, VSB, varies between −0.1 and 0.3 m s−1 and is mostly positive (northward directed, Fig. 2d). Diurnal (DU, from 33−1 to 16−1 cph) baroclinic velocities are significant in this region. Following W20, a complex EOF derived cross-shore (cross-isobath) surface diurnal velocity

Time series of (a) sea surface level η at SB, (b) 12-hourly wind vectors at SB, (c) subtidal top-to-bottom buoyancy frequency N2 at SB, and (d) subtidal depth-averaged alongshore velocity at SB VSB and surface cross-shore first-mode diurnal velocity
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
c. Surface density front identification
Surface density fronts frequently occur during the analysis period, as shown in the two examples (Fig. 3). In the first example, the ≈6-km-long surface density front is steeply angled relative to the cross-shore direction and is mostly onshore of the 25 m isobath (Fig. 3a). This front was compressed in the alongshore direction by the convergent alongshore flow (as described in W20, Fig. 3 therein). In the second example, the surface density front is much more aligned in the cross-shore direction with a length of ≈5 km (Fig. 3b).

Surface density perturbation (after removing the spatial mean within the front study region, color shading), the detected front (bold black line), and the frontal control volume (red rectangle) of (a) an inclined and (b) a cross-shore oriented front. (c) Zoom-in of the cross-shore front density perturbation ρ′ (after removing the cross-front mean) and the surface current perturbation (after removing the cross-front mean, vectors) within the control volume in (b). In (a), the cyan line denotes the shoreline normal direction. In each panel, the green dot shows the mean front location, the dashed magenta line is the front axis, and the thin black contour denotes isobaths h = [10, 25] m.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
Both example density fronts are primarily located within the front study region (see white box in Fig. 1), a bounded region (5.5 × 18.5 km2) that extends from the shoreline to the ≈30-m isobath and spans the surfzone through midshelf. The region’s southern and northern boundaries are 5 km away from the grid’s southern open boundary and 7 km from the SDB mouth, respectively. Surface density fronts primarily contained within this study region are the focus of this work.
Modeled surface density fronts are identified by applying the Canny edge detection algorithm (Canny 1986) to the surface density. This algorithm has been successfully applied to front detection in SST satellite images (e.g., Castelao et al. 2006; Jones et al. 2012). The algorithm first interpolates the surface density onto an equally spaced horizontal grid with Δ = 40-m resolution, coarser than the grid mean resolution (≈30 m) to preserve data quality. Then, density is smoothed using a 2D Gaussian filter with a
For our analysis, we focus on fronts that are relatively straight, are longer than 4 km, and are not strongly affected by open boundaries, SDB outflow, or the surfzone. Thus, we apply the following criteria to reject fronts identified by the edge detection algorithm. First, the mean front location (i.e., center of mass of the front, green dot in Figs. 3a and 3b) must be located within the front study region. Second, the offshore end of the front must be at least 1.5 km from the shoreline, to ensure that surfzone processes are not dominating the front. Third, the front is fit to an ellipse. To ensure relatively straight fronts, we require that the ratio of the ellipse minor to major axes γ < 0.15. Fourth, we require that the front length is >4 km. Both example fronts (Fig. 3) pass the criteria as their mean location is within the front study region, their lengths are >4 km, and their γ = 0.06 and γ = 0.04.
Applying the edge detection algorithm and four criteria, the total number of detected fronts Nf over the analysis period (2112 h) is a function of the threshold |∇Hρ|c (Fig. 4). As |∇Hρ|c increases from 0.2 to 12.3 × 10−4 kg m−4, the total count decreases from Nf = 6742 to Nf = 371 (Fig. 4), and the mean hourly front count decreases from 3.2 to 0.17. For the following analysis, we choose the |∇Hρ|c threshold as the inflection of the curve (triangle in Fig. 4, |∇Hρ|c = 2.9 × 10−4 kg m−4). This choice requires fronts to have a relatively strong density gradient while it allows sufficient fronts (Nf = 2948) for statistical analyses. The lower cutoff c|∇Hρ|c = 1.2 × 10−4 kg m−4 is comparable to the upper end of the smoothed 25-m isobath rms alongshore density gradient 1.5 × 10−4 kg m−4 (W20).

Total front count Nf versus the cutoff surface density gradient |∇Hρ|c. The triangle highlights the inflection point of the curve that corresponds to |∇Hρ|c = 2.9 × 10−4 kg m−4.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
For the Nf = 2948 selected fronts, kinematic front parameters are defined. First, a front axis is defined as the least squares fit line to the front (see magenta dashed line in Figs. 3a and 3b). A front orientation angle θf (θf ∈ [−90°, 90°]) is the angle between the mean shoreline normal direction (5° clockwise from the grid cross-shore orientation, see cyan line in Fig. 3a) and the front axis (see Fig. 3a). For fronts that tilt northward offshore, θf < 0° (Fig. 3a). The front length Lf is defined as the length of the front projected onto the front axis (Fig. 3b). The alongfront mean surface density gradient |∇Hρ|f is calculated by averaging the surface |∇Hρ| along the bending front. Note that |∇Hρ|f magnitude must be ≥c|∇Hρ|c. For reference, the first example front (Fig. 3a) has θf = −54°, Lf = 5.9 km, and |∇Hρ|f = 3.0 × 10−4 kg m−4 and the second example front (Fig. 3b) has θf = 6°, Lf = 5.5 km, and |∇Hρ|f = 5.6 × 10−4 kg m−4.
3. Kinematic frontal properties
a. Kinematic front parameter statistics
Here, a statistical analysis on the kinematic front parameters (θf, Lf, |∇Hρ|f, and front mean location) is performed on the Nf = 2948 selected fronts. The front orientation angle θf histogram has a U-shaped distribution (Fig. 5a) with maxima near ±90° (fronts aligned with the shoreline) and a minimum near θf = 0° (fronts shore-normal oriented). The prevalence of alongshore-oriented fronts is generally consistent with modeled inner to midshelf density gradients preferentially aligned across isobath (Romero et al. 2013; Dauhajre et al. 2017). Based on the θf distribution, fronts are categorized into alongshore oriented, cross-shore oriented, and inclined fronts. Alongshore fronts are defined as having a near-shoreline (within 20°) orientation, that is θf ∈ [−90°, −70°] or θf ∈ [70°, 90°] (dark gray shading in Fig. 5a). Cross-shore fronts are defined as having an orientation θf ∈ [−50°, 50°] (light gray shading in Fig. 5a, an example in Fig. 3b). Separating the alongshore and cross-shore fronts are inclined fronts, defined as having an orientation θf ∈ [−70°, −50°] or θf ∈ [50°, 70°] (see example in Fig. 3a). Overall, the alongshore-oriented, inclined, and cross-shore-oriented fronts account for 55%, 27%, and 18% of the total fronts, respectively. Thus, the cross-shore fronts are about 1/3 as numerous as alongshore fronts. The cross-shore fronts have the widest angular range as the cross-shore shear of the alongshore flow can tilt cross-shore fronts. Cross-shore fronts tilting northward offshore (θf < 0°) are more likely than those tiling southward offshore (θf > 0°, Fig. 5a), as the alongshore flow is mostly northward directed (see VSB in Fig. 2d).

(a) The front orientation θf histogram (blue bars), and histogram (on logarithmic scale) of (b) the front length Lf and (c) the alongfront averaged density gradient |∇Hρ|f for all the fronts (circle). The fronts are categorized into alongshore [dark gray shading in (a), star marker in (b) and (c)], inclined, and cross-shore [light gray shading in (a), triangle marker in (b) and (c)]. In (a), red histogram indicates the 86 cross-shore fronts used to create the ensemble mean front (section 4). The dashed line in (c) denotes |∇Hρ|c.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
For all fronts, the front length Lf histogram is quasi-exponential with Lf = 4 km most likely and the Lf ≥ 16 km likelihood reduced by factor of 50 (Fig. 5b). The Lf histogram for alongshore and cross-shore fronts separately is also quasi-exponential. Alongshore fronts are generally longer than cross-shore fronts. Alongshore fronts have mean (±std) Lf = 7.8 (±3.4) km, while cross-shore fronts have mean (±std) Lf = 5.8 (±1.8) km. Longer cross-shore fronts (Lf > 8 km) are more likely for more negative θf ≈ −50°. The grid offshore boundary is ≈10 km from the shoreline and the grid alongshore dimension is 36 km, possibly limiting cross-shore and alongshore front Lf.
For all fronts, the alongfront averaged density gradient |∇Hρ|f histogram is skewed with maxima at |∇Hρ|c and an exponential decrease for larger |∇Hρ|f (Fig. 5c). The |∇Hρ|f can vary by a factor of 10, from 2 to 20 × 10−4 kg m−4. Although the alongshore fronts are more numerous, both alongshore and cross-shore fronts have similar mean |∇Hρ|f with values of 4.2 × 10−4 kg m−4 and 3.9 × 10−4 kg m−4, respectively. For both cross-shore and alongshore fronts, no relationship between |∇Hρ|f and θf is evident (not shown). Two-thirds of the alongshore fronts have a positive mean cross-front density gradient (denser water onshore). For the cross-shore fronts, 90% have negative mean density gradient (lighter water to the north). The |∇Hρ|f of cross-shore fronts are ≈2 times larger than the rms alongshore density gradient along a smoothed 25 m depth contour (W20).
Next, the spatial distribution of the cross- and alongshore fronts is examined for preferred frontal position within the front study region. For example, the intermittent TJRE discharge and the TJRE shoal may promote local frontogenesis. The Nf = 528 cross-shore fronts are present throughout the front study region, mostly tilting northward offshore (Fig. 6a1). The range of Lf and ∇Hρ for the cross-shore fronts is also evident. The cross-shore front mean location (center of mass) has ≈2/3 of fronts located northward of the TJRE mouth, and ≈1/3 located south of the TJRE mouth (Fig. 6a2). The alongshore fronts also are present throughout the front study region (Fig. 6b1). Alongshore front mean cross-shore location is twice as likely to occur at the midpoint of the front study region rather than its offshore end (Fig. 6b2). Alongshore front mean location is somewhat more likely found south of the TJRE mouth, relative to the north.

(a1),(b1) Front spatial distribution and (a2),(b2) the binned mean front location for all (a) cross-shore and (b) alongshore fronts. In (a1) and (b1), the front color represents the alongfront averaged density gradient |∇Hρ|f. Black contours in (a2) and (b2) denote the 10- and 25-m isobaths.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
b. Cross-shore and alongshore front occurrence frequency
The differences in front kinematic parameters (θf, Lf, center of mass, Figs. 5 and 6) suggest that different processes are responsible for generating the cross-shore and alongshore fronts. Here, we examine the factors affecting the temporal variability of frontal occurrences and the mean |∇Hρ|f for both cross-shore and alongshore fronts. The hourly front count nf(t) is defined as the number of identified fronts for a particular hour. For cross-shore fronts, the hourly front count nf varies between 0 and 5, with a time mean (±std) of 0.25 (±0.60) (Fig. 7b). Cross-shore front hourly nf are elevated during four time periods (i.e., 2–9 August, 29–30 August, 8–16 September, and 12–16 October), coincident with the periods of positive (northward) VSB (Fig. 7a). The hourly nf of the alongshore fronts ranges between 0 and 10, with a time mean (± std) of 0.75 (±1.2) (Fig. 7c). The alongshore front nf is not related to VSB. Alongshore fronts are detected for 84 out of 88 days of the analysis period (i.e., 95% of the period), and the nf has consistent diurnal variability.

Time series of (a) subtidal alongshore velocity VSB (black) and the diurnal-band surface baroclinic cross-shore velocity
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
To further investigate the relationship between frontal temporal variability and flow conditions, the cross-shore and alongshore hourly nf are bin averaged using VSB and the surface diurnal velocity

(a) Bin-averaged hourly front count nf for the cross-shore fronts and the standard error (error bar) vs the subtidal alongshore velocity at SB VSB. (b) Bin-averaged hourly nf for the alongshore fronts and the standard error (error bar) vs the diurnal-band first-mode surface cross-shore velocity
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
For the alongshore fronts, the binned-mean hourly nf is small (≤0.5) for
4. Ensemble mean cross-shore front
Individual cross-shore fronts have variable orientation θf, frontal length Lf, and frontal density gradient |∇Hρ|f. These fronts can be slightly curved, and the front may deviate from the front axis. To better understand the characteristics and dynamics of the cross-shore fronts, an ensemble mean cross-shore front is created in the following analysis. Analysis of alongshore fronts will be further investigated elsewhere.
a. Cross-shore front extraction, front decomposition, and ensemble average
To diagnose ensemble cross-shore front dynamics, individual cross-shore fronts are first extracted, variables are decomposed into cross-front mean and perturbation components, and an ensemble mean cross-shore front is then generated on a subset of the cross-shore fronts. Each cross-shore front is extracted using a rectangular control volume, centered along the best-fit front axis, with horizontal dimensions of 4 km cross-front and 8 km alongfront where the onshore end of the control volume intersects the shoreline. Control volumes for the two example fronts are shown in Figs. 3a and 3b. Within the control volume, an alongfront coordinate (
The extracted cross-shore fronts (Nf = 528, Fig. 6a) have a wide range of θf. Some cross-shore fronts may interact with adjacent fronts detected at the same time. Most but not all cross-shore fronts have a negative cross-front density gradient
- For a particular cross-shore front, all other fronts detected at the same time step must be separated by >4 km (the control volume width). This criterion reduces the total cross-shore front count from 528 to 431.
- Cross-shore fronts should be roughly shore-normal requiring the front orientation angle (θf ∈ [−25°, 25°], removing an additional 299 fronts. A wider θf range (e.g., ±45°) obtains consistent results for the subsequent analyses (not shown here).
- The front must reach the 25-m isobath (see Fig. 3b) and must span
within the control volume (black contour in Fig. 3c), ensuring the front spans across the inner to midshelf, and allowing an alongfront average over this region. This criterion excludes an additional 42 fronts. - To ensure consistent ensemble front dynamics, the cross-front density gradient
must be negative, excluding four remaining fronts with positive . These four fronts are associated with the northern side of cross-shore-oriented warm filaments that occur infrequently.
These criteria together result in a subset of 86 cross-shore test fronts (Fig. 9). The example cross-shore front in Fig. 3b is also a test front. The test fronts mostly tilt gently northward offshore, and their θf (red bar in Fig. 5a) are mostly negative with mean (±std) of −10.5° (±12.3°). The test fronts are concentrated north of the TJRE mouth (Fig. 9). The test fronts have a mean Lf = 5.9 km, similar to mean Lf = 5.8 km for all cross-shore fronts. In addition, the test fronts have a mean |∇Hρ|f = 3.5 × 10−4 kg m−4, also similar to the mean of 3.9 × 10−4 kg m−4 for all cross-shore fronts (section 3a). This suggests that the 86 test fronts are representative of the set of all cross-shore fronts. Consistent with the entire set of cross-shore fronts, test fronts occur (red line in Fig. 7b) only for VSB ≥ 0.08 m s−1 and also have stronger density gradient for VSB > 0.16 m s−1 (not shown). Among the test fronts, 56 fronts are from 17 fronts that are extracted at multiple (2–6) times as the front is advected. The other 30 fronts are unique and extracted only once. Overall, 47 unique fronts are included.

Spatial distribution of the 86 cross-shore test fronts used to create the ensemble mean front with color representing |∇Hρ|f. Black contours denote the 10- and 25-m isobaths.
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
The ensemble average [Eq. (3)] is estimated at a particular
Recall that all test fronts are required to have a front within the range of
b. Ensemble and cross-front averaged test front
The ensemble and cross-front mean variables represent the ensemble background shelf conditions associated with the test fronts (Fig. 10). The ensemble and cross-front mean sea surface elevation

(a) Ensemble and cross-front averaged sea surface elevation
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
The ensemble mean of the cross-front perturbation variables reveal a clear front that has dense (

Plan view of (a) the ensemble perturbation sea surface elevation ⟨η′⟩, and the near surface (at z = −1 m) ensemble perturbation (b) temperature ⟨T′⟩, (c) density ⟨ρ′⟩, (d) alongfront velocity ⟨u′⟩, (e) cross-front velocity ⟨υ′⟩, and (f) vertical velocity ⟨w′⟩. The blue dashed lines delineate the region (
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
To analyze the cross-front and vertical structure of the ensemble front, we additionally alongfront average ensemble-mean perturbation variation within

Ensemble and alongfront (
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
The ensemble mean perturbation vertical vorticity
The ensemble averaging of the 86 test fronts with variable lengths, density gradients, and deviation from a straight line results in some smoothing of the resulting ensemble front. We evaluate the smoothing by first examining the coincident parameters of the example front in Figs. 3b and 3c. The example front density is also temperature dominated and has maximum density gradient
5. Ensemble front frontogenesis tendency

Ensemble and alongfront (
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
The cross-front shear induced
6. Ensemble front momentum balance
For the perturbation velocity, no alongfront jet develops at the front axis (Fig. 12d), different from both the DF (Hoskins and Bretherton 1972) and TTW (McWilliams 2017) mechanisms, where an alongfront jet is in approximate thermal wind balance. The frontogenesis tendency analysis (Fig. 13) indicates the involvement of ageostrophic processes. In the DF and TTW mechanisms, an ageostrophic secondary cross-front flow υa is induced and the ageostrophic Coriolis forcing fυa is balanced by the alongfront material acceleration in the DF mechanism (Hoskins et al. 1978), and the vertical mixing in the TTW mechanism (McWilliams et al. 2015). Here the along-/cross-front momentum balances are examined and compared with these mechanisms.

Ensemble and alongfront (
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1

As in Fig. 14, but for the perturbation momentum terms in the cross-front (
Citation: Journal of Physical Oceanography 51, 3; 10.1175/JPO-D-20-0162.1
For the alongfront momentum terms in Eq. (7a), the perturbation Coriolis forcing
For the cross-front momentum terms, the perturbation
Overall, the along-ensemble front ageostrophic balance is analogous to that in the DF mechanism (Hoskins et al. 1978; Thomas et al. 2008). One striking difference is that the ensemble front is not in cross-front geostrophic balance, whereas the DF mechanism has an approximate cross-front geostrophic balance (e.g., Hoskins and Bretherton 1972). The cross-front momentum balance between
7. Discussion
a. Coastal density front properties
Within the 3-month analysis period, the density gradient magnitude and orientation contrast with previous studies. Here, the number of cross-shore fronts (with a ±50° range of θf), is one-third that of alongshore fronts (Figs. 5a and 6). Even with taking the θf range into account, this contrasts with previous high (75 m) resolution coastal numerical model results (Dauhajre et al. 2017), where a cross-isobath density gradient was a factor 20 more probable than an along-isobath density gradient in depths ≤ 50 m. Here, the subtidal stratification N2 is relatively strong from 10−4 to 4 × 10−4 s−2 (Fig. 2c), consistent with regional observations (e.g., Palacios et al. 2004). Although Dauhajre et al. (2017) do not report N2, sections through springtime fronts and filaments allow inference of N2 ≈ 10−5 s−2, an order of magnitude weaker than the San Diego Bight simulation. The ensemble cross-shore front has cross-front density gradient
The horizontal density gradients here also are larger than deeper-water frontal horizontal density gradients of O(10−4)–O(10−5) observed in the CCS (Pallàs-Sanz et al. 2010; Johnson et al. 2020) over scales of 2–5 km. These differences may be due to the background LV4 meridional density gradient, the general shoreward strengthening of surface horizontal density gradients from deep (>500 m) water to the shelf (e.g., Dauhajre et al. 2017), and the high (~30 m) grid resolution relative to O(1) km in observations.
Divergence and vorticity are also key front parameters. The ensemble cross-shore front has maximum divergence magnitude
Spatial distribution of the Nf = 528 cross-shore fronts shows a concentration (≈2/3 of the fronts) to the north of the TJRE mouth (Fig. 6a2), suggesting that the TJRE shoal may be a factor in promoting cross-shore front generation, as VSB is mostly positive (northward directed). The alongshore front occurrence frequency is elevated with onshore directed surface diurnal baroclinic flow
b. Comparison with the TTW and DF mechanisms
The TTW mechanism has been invoked to explain density filament generation in specific case studies (e.g., Gula et al. 2014; Dauhajre et al. 2017). A case study of two Gulf Stream density filaments showed that ageostrophic Coriolis forcing is balanced by the vertical mixing (Gula et al. 2014). Another case study of two density filaments and fronts in 20–30-m water depth on the shelf with wind stress ~0.03 N m−2 found that the horizontal flow field is consistent with the TTW dynamics (Dauhajre et al. 2017). This wind stress was roughly a factor of 3 times stronger than the typical subtidal wind stress ~0.01 N m−2 in the San Diego Bight simulations, implying wind speeds 1.7 times stronger, consistent with regional spring to fall differences (e.g., Winant and Dorman 1997; Dong et al. 2009). These case studies analyzed individual hand-selected filaments, in contrast to the ensemble front that comprises 86 test fronts.
In the DF mechanism, the large-scale strain field is nondivergent and in geostrophic balance (e.g., Hoskins and Bretherton 1972; McWilliams 2017). Here,
This difference in DF-mechanism ASC and ensemble cross-shore front perturbation momentum balances may be due to the presence of a shoreline boundary. Coastal circulation also is largely semigeostrophic (e.g., Allen 1980; Lentz et al. 1999), particularly at subtidal time scales, with largely geostrophic cross-shore momentum balance and largely ageostrophic alongshore balance as wind forcing and nonlinear advection can become important. For open ocean fronts, the alongfront flow is unbounded and the cross-front momentum balance can be consistent with a geostrophic (thermal wind) balance. Here the ensemble cross-shore front is constrained by the shoreline and the associated semigeostrophic momentum balance becomes more consistent with that of coastal circulation. In the end, cross-front convergence is key across various types of surface density fronts from the unbounded TTW and DF to the shoreline-bounded ensemble cross-shore front here, albeit via different dynamics.
c. The cross-front ageostrophic balance: Relationship to a gravity current
The primary cross-front momentum balance between the perturbation
The ensemble cross-shore front has similarities with coastal buoyant plume fronts that typically have order of magnitude larger density gradients (e.g., Lentz et al. 2003). Both are cross-shore-oriented, shoreline bounded, and have a cross-front ageostrophic balance. Modeled coastal buoyant plumes with much larger density gradients have gravity current dynamics (e.g., Akan et al. 2018). In Lentz et al. (2003), the observed plume front propagates alongshore from the light toward the dense side with a cross-front density difference Δρ ≈ 3.0 kg m−3 over 2 km, and propagation speed reaching ≈0.5 m s−1. For the ensemble front, cross-front density difference Δ⟨ρ′⟩ ≈ −0.08 kg m−3 over 1 km (Fig. 11c), 37 times weaker than the Lentz et al. (2003) plume front. Assuming that the ensemble front represents a two-layer gravity current in h = 20-m depth with an upper (lower) layer depth of h1 = 5 m (h2 = 15 m), the corresponding gravity current speed is
We next explore why the ensemble cross-shore front (made up of the 86 test fronts) has dynamics similar to a gravity current. Modeled density fronts in geostrophic balance can transform into gravity currents for similar density differences as seen here (Warner et al. 2018; Pham and Sarkar 2018). However, the shoreline boundary constrains the cross-shore flow preventing cross-shore-oriented fronts from being in near-geostrophic balance, as suggested by the cross-front momentum balance. Note that alongshore-oriented fronts, such as the example close to a headland in Dauhajre et al. (2017), have no such limitation. Near-field river plumes behave like gravity currents but for distances larger than a Rossby deformation radius LR = Nh/f other dynamics are important. The cross-shore fronts are 7–18 km from the SDB mouth, and using N2 = 2 × 10−4 s−2 and h = 25 m results in LR ≈ 4.5 km. This and the lack of relationship between cross-shore frontal occurrence and SDB outflow indicate that the cross-shore fronts are not gravity currents directly forced by the SDB outflow.
The modeled San Diego Bight region has a weak (a factor of 30 times smaller than the ensemble cross-shore front) large-scale alongshore density gradient (W20) due to regional upwelling gradients (e.g., Huyer 1983) and warm water outflow from the SDB. W20 noted that the northward directed subtidal depth-averaged alongshore flow along the ≈25-m isobath was convergent with divergence of ≈−0.05f, much weaker than the divergence of the ensemble front. However, the divergent northward flow acting on the large-scale density gradient was found to enhance root-mean-square alongshore density gradients (W20). Here, the cross-shore front occurrence and elevated density gradients were much more likely for stronger northward subtidal flow (Fig. 8a). This suggests that the cross-shore fronts, whose ensemble had gravity current like dynamics, is generated by the combined convergent northward flow acting on the large-scale density field.
8. Summary
Here, we investigate the kinematics and dynamics of the coastal (within 10 km from shore and <30-m water depth) density fronts, using a high-resolution numerical model of the San Diego Bight (W20). Density fronts are first identified using the Canny edge detection algorithm and then categorized into alongshore and cross-shore-oriented fronts. Statistics of front properties show that, the cross-shore fronts are about 1/3 as numerous as the alongshore fronts. For both front groups, the mean front length reaches 6–8 km, the alongfront averaged surface density gradient varies from 2 to 20 × 10−4 kg m−4. Most (≈2/3) alongshore fronts have lighter water offshore, while 90% of cross-shore fronts have lighter water to the north. The alongshore front activity is enhanced by onshore surface diurnal flow, indicating onshore propagating internal warm bores. In contrast, the cross-shore front activity is promoted by northward subtidal alongshore flow.
The cross-shore front dynamics are further examined using a subset of the cross-shore fronts that have a negative cross-front density gradient (lighter water to the north). The density and flow field are decomposed into cross-front mean and perturbation components, and then ensemble averaged to generate an ensemble cross-shore front. The cross-front mean flow is largely in geostrophic balance in the along- and cross-front directions. The ensemble front extends several kilometers from shore with a distinct linear front axis and convergent perturbation cross-front flow within the upper 5 m. The perturbation alongfront flow within the upper 5 m is more offshore (onshore) directed on the light (dense) side and weakens onshore. Downwelling occurs on the front dense side, and weaker upwelling occurs on the light side. The ensemble mean front is frontogenetic as the cross-front convergence dominates over the frontolytic vertical advection. Vertical mixing of momentum is weak, indicating that the turbulent thermal wind mechanism is not active. The perturbation alongfront momentum balance is largely geostrophic, while the cross-front balance is between the pressure gradient and the material acceleration, analogous to a gravity current. This contrasts with the cross-front geostrophic and alongfront ageostrophic balances in classic deformation frontognesis, but is consistent with shoreline-bounded semigeostrophic coastal circulation. Given that alongshore nonuniform density and alongshore convergent flows are ubiquitous in coastal waters, shallow cross-shore fronts may also occur at many other locations.
This work was supported by the National Science Foundation (OCE-1459389) as part of the Cross-Surfzone/Inner-shelf Dye Exchange (CSIDE) experiment. Additional funding is through the Environmental Protection Agency through the North American Development Bank, however it does not necessarily reflect the policies, actions, or positions of the U.S. EPA or NADB. This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation (ACI-1548562). The numerical simulations were performed on the comet cluster at the San Diego Supercomputer Center through XSEDE allocation TG-OCE180013. NOAA provided the NAM atmospheric forcing fields and the bathymetry. SIO Coastal Data Information Program provided wave forcing. Ganesh Gopalakrishnan and Bruce Cornuelle provided CASE model solutions which are available online (http://ecco.ucsd.edu/case.html). We also appreciate extra support from the Tijuana River National Estuarine Research Reserve and the Southern California Coastal Ocean Observing System. Geno Pawlak, Derek Grimes, Angelica Rodriguez, and Nirnimesh Kumar provided useful feedback on this work. We thank two anonymous reviewers for helpful comments that improved this manuscript.
REFERENCES
Akan, Ç., J. C. McWilliams, S. Moghimi, and H. T. Özkan-Haller, 2018: Frontal dynamics at the edge of the Columbia River plume. Ocean Modell., 122, 1–12, https://doi.org/10.1016/j.ocemod.2017.12.001.
Allen, J. S., 1980: Models of wind-driven currents on the continental shelf. Annu. Rev. Fluid Mech., 12, 389–433, https://doi.org/10.1146/annurev.fl.12.010180.002133.
Austin, J. A., and J. A. Barth, 2002: Variation in the position of the upwelling front on the Oregon shelf. J. Geophys. Res., 107, 3180, https://doi.org/10.1029/2001JC000858.
Austin, J. A., and S. J. Lentz, 2002: The inner shelf response to wind-driven upwelling and downwelling. J. Phys. Oceanogr., 32, 2171–2193, https://doi.org/10.1175/1520-0485(2002)032<2171:TISRTW>2.0.CO;2.
Banas, N., P. MacCready, and B. Hickey, 2009: The columbia river plume as cross-shelf exporter and along-coast barrier. Cont. Shelf Res., 29, 292–301, https://doi.org/10.1016/j.csr.2008.03.011.
Barkan, R., M. J. Molemaker, K. Srinivasan, J. C. McWilliams, and E. A. D. Asaro, 2019: The role of horizontal divergence in submesoscale frontogenesis. J. Phys. Oceanogr., 49, 1593–1618, https://doi.org/10.1175/JPO-D-18-0162.1.
Benjamin, T. B., 1968: Gravity currents and related phenomena. J. Fluid Mech., 31, 209–248, https://doi.org/10.1017/S0022112068000133.
Bleck, R., R. Onken, and J. Woods, 1988: A two-dimensional model of mesoscale frontogenesis in the ocean. Quart. J. Roy. Meteor. Soc., 114, 347–371, https://doi.org/10.1002/qj.49711448005.
Booij, N., R. C. Ris, and L. H. Holthuijsen, 1999: A third-generation wave model for coastal regions: 1. Model description and validation. J. Geophys. Res., 104, 7649–7666, https://doi.org/10.1029/98JC02622.
Brink, K., 1987: Upwelling fronts: Implications and unknowns. S. Afr. J. Mar. Sci., 5, 3–9, https://doi.org/10.2989/025776187784522315.
Callies, J., R. Ferrari, J. M. Klymak, and J. Gula, 2015: Seasonality in submesoscale turbulence. Nat. Commun., 6, 6862, https://doi.org/10.1038/ncomms7862.
Canny, J., 1986: A computational approach to edge detection. IEEE Trans. Pattern Anal. Mach. Intell., PAMI-8, 679–698, https://doi.org/10.1109/TPAMI.1986.4767851.
Castelao, R. M., T. P. Mavor, J. A. Barth, and L. C. Breaker, 2006: Sea surface temperature fronts in the California Current System from geostationary satellite observations. J. Geophys. Res., 111, C09026, https://doi.org/10.1029/2006JC003541.
Chant, R., 2011: Interactions between estuaries and coasts: River plumes their formation, transport, and dispersal. Treatise on Estuarine and Coastal Science, E. Wolanski and D. McLusky, Eds., Academic Press, 213–235, https://doi.org/10.1016/B978-0-12-374711-2.00209-6.
Colosi, J. A., N. Kumar, S. H. Suanda, T. M. Freismuth, and J. H. MacMahan, 2018: Statistics of internal tide bores and internal solitary waves observed on the inner continental shelf off Point Sal, California. J. Phys. Oceanogr., 48, 123–143, https://doi.org/10.1175/JPO-D-17-0045.1.
Connolly, T. P., and A. R. Kirincich, 2019: High-resolution observations of subsurface fronts and alongshore bottom temperature variability over the inner shelf. J. Geophys. Res. Oceans, 124, 593–614, https://doi.org/10.1029/2018JC014454.
Dauhajre, D. P., J. C. McWilliams, and Y. Uchiyama, 2017: Submesoscale coherent structures on the continental shelf. J. Phys. Oceanogr., 47, 2949–2976, https://doi.org/10.1175/JPO-D-16-0270.1.
Dauhajre, D. P., J. C. McWilliams, and L. Renault, 2019: Nearshore Lagrangian connectivity: Submesoscale influence and resolution sensitivity. J. Geophys. Res. Oceans, 124, 5180–5204, https://doi.org/10.1029/2019JC014943.
Dewey, R. K., and J. N. Moum, 1990: Enhancement of fronts by vertical mixing. J. Geophys. Res., 95, 9433–9445, https://doi.org/10.1029/JC095iC06p09433.
Dinniman, M. S., and M. M. Rienecker, 1999: Frontogenesis in the North Pacific oceanic frontal zones—A numerical simulation. J. Phys. Oceanogr., 29, 537–559, https://doi.org/10.1175/1520-0485(1999)029<0537:FITNPO>2.0.CO;2.
Dong, C., E. Y. Idica, and J. C. McWilliams, 2009: Circulation and multiple-scale variability in the Southern California Bight. Prog. Oceanogr., 82, 168–190, https://doi.org/10.1016/j.pocean.2009.07.005.
Feddersen, F., M. Olabarrieta, R. T. Guza, D. Winters, B. Raubenheimer, and S. Elgar, 2016: Observations and modeling of a tidal inlet dye tracer plume. J. Geophys. Res. Oceans, 121, 7819–7844, https://doi.org/10.1002/2016JC011922.
Franks, P. J., 1992: Phytoplankton blooms at fronts: Patterns, scales, and physical forcing mechanisms. Rev. Aquat. Sci., 6, 121–137.
Garrett, C. J. R., and J. Loder, 1981: Dynamical aspects of shallow sea fronts. Philos. Trans. Roy. Soc. London, 302A, 563–581, https://doi.org/10.1098/rsta.1981.0183.
Geyer, W. R., and D. K. Ralston, 2015: Estuarine frontogenesis. J. Phys. Oceanogr., 45, 546–561, https://doi.org/10.1175/JPO-D-14-0082.1.
Grimes, D. J., F. Feddersen, S. N. Giddings, and G. Pawlak, 2020: Cross-shore deformation of a surfzone-released dye plume by an internal tide on the inner shelf. J. Phys. Oceanogr., 50, 35–54, https://doi.org/10.1175/JPO-D-19-0046.1.
Gula, J., M. J. Molemaker, and J. C. McWilliams, 2014: Submesoscale cold filaments in the Gulf Stream. J. Phys. Oceanogr., 44, 2617–2643, https://doi.org/10.1175/JPO-D-14-0029.1.
Hally-Rosendahl, K., F. Feddersen, D. B. Clark, and R. Guza, 2015: Surfzone to inner-shelf exchange estimated from dye tracer balances. J. Geophys. Res. Oceans, 120, 6289–6308, https://doi.org/10.1002/2015JC010844.
Horner-Devine, A. R., R. D. Hetland, and D. G. MacDonald, 2015: Mixing and transport in coastal river plumes. Annu. Rev. Fluid Mech., 47, 569–594, https://doi.org/10.1146/annurev-fluid-010313-141408.
Hoskins, B. J., 1982: The mathematical theory of frontogenesis. Annu. Rev. Fluid Mech., 14, 131–151, https://doi.org/10.1146/annurev.fl.14.010182.001023.
Hoskins, B. J., and F. P. Bretherton, 1972: Atmospheric frontogenesis models: Mathematical formulation and solution. J. Atmos. Sci., 29, 11–37, https://doi.org/10.1175/1520-0469(1972)029<0011:AFMMFA>2.0.CO;2.
Hoskins, B. J., I. Draghici, and H. Davies, 1978: A new look at the ω-equation. Quart. J. Roy. Meteor. Soc., 104, 31–38, https://doi.org/10.1002/qj.49710443903.
Huyer, A., 1983: Coastal upwelling in the California Current System. Prog. Oceanogr., 12, 259–284, https://doi.org/10.1016/0079-6611(83)90010-1.
Huzzey, L. M., and J. M. Brubaker, 1988: The formation of longitudinal fronts in a coastal plain estuary. J. Geophys. Res., 93, 1329–1334, https://doi.org/10.1029/JC093iC02p01329.
Johnson, E. S., 1996: A convergent instability wave front in the central tropical Pacific. Deep-Sea Res. II, 43, 753–778, https://doi.org/10.1016/0967-0645(96)00034-3.
Johnson, L., C. M. Lee, E. A. D. Asaro, L. Thomas, and A. Shcherbina, 2020: Restratification at a California Current upwelling front. Part I: Observations. J. Phys. Oceanogr., 50, 1455–1472, https://doi.org/10.1175/JPO-D-19-0203.1.
Jones, C. T., T. D. Sikora, P. W. Vachon, and J. Wolfe, 2012: Toward automated identification of sea surface temperature front signatures in Radarsat-2 images. J. Atmos. Oceanic Technol., 29, 89–102, https://doi.org/10.1175/JTECH-D-11-00088.1.
Kahru, M., E. Di Lorenzo, M. Manzano-Sarabia, and B. G. Mitchell, 2012: Spatial and temporal statistics of sea surface temperature and chlorophyll fronts in the California Current. J. Plankton Res., 34, 749–760, https://doi.org/10.1093/plankt/fbs010.
Kantha, L. H., and C. A. Clayson, 1994: An improved mixed layer model for geophysical applications. J. Geophys. Res., 99, 25 235–25 266, https://doi.org/10.1029/94JC02257.
Kumar, N., G. Voulgaris, J. C. Warner, and M. Olabarrieta, 2012: Implementation of the vortex force formalism in the Coupled Ocean-Atmosphere-Wave-Sediment Transport (COAWST) modeling system for inner shelf and surf zone applications. Ocean Modell., 47, 65–95, https://doi.org/10.1016/j.ocemod.2012.01.003.
Kumar, N., F. Feddersen, Y. Uchiyama, J. McWilliams, and W. O. Reilly, 2015: Midshelf to surfzone coupled ROMS–SWAN model data comparison of waves, currents, and temperature: Diagnosis of subtidal forcings and response. J. Phys. Oceanogr., 45, 1464–1490, https://doi.org/10.1175/JPO-D-14-0151.1.
Kumar, N., S. H. Suanda, J. A. Colosi, K. Haas, E. Di Lorenzo, A. J. Miller, and C. A. Edwards, 2019: Coastal semidiurnal internal tidal incoherence in the Santa Maria basin, California: Observations and model simulations. J. Geophys. Res. Oceans, 124, 5158–5179, https://doi.org/10.1029/2018JC014891.
Lentz, S. J., R. T. Guza, S. Elgar, F. Feddersen, and T. H. C. Herbers, 1999: Momentum balances on the North Carolina inner shelf. J. Geophys. Res., 104, 18 205–18 226, https://doi.org/10.1029/1999JC900101.
Lentz, S. J., S. Elgar, and R. Guza, 2003: Observations of the flow field near the nose of a buoyant coastal current. J. Phys. Oceanogr., 33, 933–943, https://doi.org/10.1175/1520-0485(2003)33<933:OOTFFN>2.0.CO;2.
Lévy, M., P. J. Franks, and K. S. Smith, 2018: The role of submesoscale currents in structuring marine ecosystems. Nat. Commun., 9, 4758, https://doi.org/10.1038/s41467-018-07059-3.
Mahadevan, A., 2016: The impact of submesoscale physics on primary productivity of plankton. Annu. Rev. Mar. Sci., 8, 161–184, https://doi.org/10.1146/annurev-marine-010814-015912.
Mauzole, Y., H. Torres, and L.-L. Fu, 2020: Patterns and dynamics of SST fronts in the California Current System. J. Geophys. Res. Oceans, 125, e2019JC015499, https://doi.org/10.1029/2019JC015499.
McSweeney, J. M., and Coauthors, 2020: Observations of shoaling nonlinear internal bores across the central California inner shelf. J. Phys. Oceanogr., 50, 111–132, https://doi.org/10.1175/JPO-D-19-0125.1.
McWilliams, J. C., 2016: Submesoscale currents in the ocean. Proc. Roy. Soc., 472A, 20160117, https://doi.org/10.1098/rspa.2016.0117.
McWilliams, J. C., 2017: Submesoscale surface fronts and filaments: Secondary circulation, buoyancy flux, and frontogenesis. J. Fluid Mech., 823, 391–432, https://doi.org/10.1017/jfm.2017.294.
McWilliams, J. C., J. Gula, M. J. Molemaker, L. Renault, and A. F. Shchepetkin, 2015: Filament frontogenesis by boundary layer turbulence. J. Phys. Oceanogr., 45, 1988–2005, https://doi.org/10.1175/JPO-D-14-0211.1.
Moum, J. N., J. M. Klymak, J. D. Nash, A. Perlin, and W. D. Smyth, 2007: Energy transport by nonlinear internal waves. J. Phys. Oceanogr., 37, 1968–1988, https://doi.org/10.1175/JPO3094.1.
Nagai, T., N. Gruber, H. Frenzel, Z. Lachkar, J. C. McWilliams, and G.-K. Plattner, 2015: Dominant role of eddies and filaments in the offshore transport of carbon and nutrients in the California Current System. J. Geophys. Res. Oceans, 120, 5318–5341, https://doi.org/10.1002/2015JC010889.
O’Donnell, J., 2010: The dynamics of estuary plumes and fronts. Contemporary Issues in Estuarine Physics, Cambridge University Press, 186–246, https://doi.org/10.1017/CBO9780511676567.009.
Ohlmann, J., M. Molemaker, B. Baschek, B. Holt, G. Marmorino, and G. Smith, 2017: Drifter observations of submesoscale flow kinematics in the coastal ocean. Geophys. Res. Lett., 44, 330–337, https://doi.org/10.1002/2016GL071537.
Palacios, D. M., S. J. Bograd, R. Mendelssohn, and F. B. Schwing, 2004: Long-term and seasonal trends in stratification in the California Current, 1950–1993. J. Geophys. Res., 109, C10016, https://doi.org/10.1029/2004JC002380.
Pallàs-Sanz, E., T. Johnston, and D. Rudnick, 2010: Frontal dynamics in a California Current System shallow front: 1. Frontal processes and tracer structure. J. Geophys. Res., 115, C12067, https://doi.org/10.1029/2009JC006032.
Pham, H. T., and S. Sarkar, 2018: Ageostrophic secondary circulation at a submesoscale front and the formation of gravity currents. J. Phys. Oceanogr., 48, 2507–2529, https://doi.org/10.1175/JPO-D-17-0271.1.
Rao, S., J. Pringle, and J. Austin, 2011: Upwelling relaxation and estuarine plumes. J. Geophys. Res., 116, C09023, https://doi.org/10.1029/2010JC006739.
Romero, L., Y. Uchiyama, J. C. Ohlmann, J. C. McWilliams, and D. A. Siegel, 2013: Simulations of nearshore particle-pair dispersion in Southern California. J. Phys. Oceanogr., 43, 1862–1879, https://doi.org/10.1175/JPO-D-13-011.1.
Romero, L., D. A. Siegel, J. C. McWilliams, Y. Uchiyama, and C. Jones, 2016: Characterizing storm water dispersion and dilution from small coastal streams. J. Geophys. Res. Oceans, 121, 3926–3943, https://doi.org/10.1002/2015JC011323.
Shchepetkin, A. F., and J. C. McWilliams, 2005: The Regional Oceanic Modeling System (ROMS): A split-explicit, free-surface, topography-following-coordinate oceanic model. Ocean Modell., 9, 347–404, https://doi.org/10.1016/j.ocemod.2004.08.002.
Simpson, J., C. Allen, and N. Morris, 1978: Fronts on the continental shelf. J. Geophys. Res., 83, 4607–4614, https://doi.org/10.1029/JC083iC09p04607.
Sinnett, G., F. Feddersen, A. J. Lucas, G. Pawlak, and E. Terrill, 2018: Observations of nonlinear internal wave run-up to the surfzone. J. Phys. Oceanogr., 48, 531–554, https://doi.org/10.1175/JPO-D-17-0210.1.
Spall, M. A., 1995: Frontogenesis, subduction, and cross-front exchange at upper ocean fronts. J. Geophys. Res., 100, 2543–2557, https://doi.org/10.1029/94JC02860.
Suanda, S. H., F. Feddersen, and N. Kumar, 2017: The effect of barotropic and baroclinic tides on coastal stratification and mixing. J. Geophys. Res. Oceans, 122, 10 156–10 173, https://doi.org/10.1002/2017JC013379.
Sutherland, B. R., 2010: Internal Gravity Waves. Cambridge University Press, 377 pp., https://doi.org/10.1017/CBO9780511780318.
Thomas, L. N., A. Tandon, and A. Mahadevan, 2008: Submesoscale processes and dynamics. Ocean Modeling in an Eddying Regime, Geophys. Monogr., Vol. 177, Amer. Geophys. Union, 17–38, https://doi.org/10.1029/177GM04.
Thompson, L., 2000: Ekman layers and two-dimensional frontogenesis in the upper ocean. J. Geophys. Res., 105, 6437–6451, https://doi.org/10.1029/1999JC900336.
Towns, J., and Coauthors, 2014: Xsede: Accelerating scientific discovery. Comput. Sci. Eng., 16, 62–74, https://doi.org/10.1109/MCSE.2014.80.
Umlauf, L., and H. Burchard, 2003: A generic length-scale equation for geophysical turbulence models. J. Mar. Res., 61, 235–265, https://doi.org/10.1357/002224003322005087.
Warner, J. C., B. Armstrong, R. He, and J. B. Zambon, 2010: Development of a Coupled Ocean-Atmosphere-Wave-Sediment Transport (COAWST) modeling system. Ocean Modell., 35, 230–244, https://doi.org/10.1016/j.ocemod.2010.07.010.
Warner, S. J., R. M. Holmes, E. H. M. Hawkins, M. S. Hoecker-Martínez, A. C. Savage, and J. N. Moum, 2018: