1. Introduction
Baroclinic instabilities are ubiquitous throughout the global ocean. They release potential energy stored in horizontal density gradients, creating unsteady and evolving motions in the flow at approximately the first baroclinic deformation radius. As such, the Burger number (Bu) associated with baroclinicly unstable flow is Bu = RdL−1 = RoRi1/2 ~ 1 (Eady 1949; Stone 1966, 1970). Beyond this, there are two general categories of baroclinic instabilities associated with the Rossby number (Ro) and the Richardson number (Ri) of the flow: mesoscale instabilities within the large-scale, geostrophic fronts with Ro ≪ 1 and Ri ≫ 1, and submesoscale instabilities with Ro ~ 1 and Ri ~ 1 (Boccaletti et al. 2007). Examples of the submesoscale instabilities include instabilities within the ocean mixed layers (Boccaletti et al. 2007; Fox-Kemper et al. 2008), along the fronts of mesoscale eddies (Callies et al. 2015; Brannigan et al. 2017), in the bottom boundary layers (Wenegrat et al. 2018), and, rarely, in certain coastal buoyancy-driven flows (Hetland 2017); they are potentially more nonlinear and energetic in nature than can be accurately described by quasigeostrophic (QG) theory.
The dynamics of QG baroclinic instabilities over a flat bottom has been well studied since the seminal papers by Eady (1949) and Phillips (1954), few studies have investigated the nongeostrophic scenario with a flat bottom (Stone 1966, 1970) and the QG scenario with a sloping bottom (Blumsack and Gierasch 1972; Mechoso 1980), but fewer studies have considered both the sloping-bathymetry effect and nongeostrophic effect. The submesoscale baroclinic instabilities in bottom boundary layers and coastal buoyancy-driven flows are influenced by these two effect (Wenegrat et al. 2018; Hetland 2017), and are less understood than the submesoscale instabilities in the surface mixed layers and the mesoscale instabilities, which are not influenced by the effects at the same time. One unique feature is the instability suppression—the submesoscale instabilities over sloping bathymetry exhibit weaker growth compared to the QG theories (Blumsack and Gierasch 1972; Mechoso 1980). Based on linear stability analysis, Wenegrat et al. (2018) shows that the growth of the instabilities in bottom boundary layers decreases as the regime shifts from QG to nongeostrophic. Also, although coastal buoyancy-driven flows are often associated with stronger lateral density gradients than open ocean fronts, they are seldom observed to be associated with instabilities; in particular, baroclinic instabilities are seldom observed in river plumes, even though lateral buoyancy gradients within the fronts are strong (Horner-Devine et al. 2015). External forcing agents, such as winds and tides, could suppress baroclinic instabilities through mixing processes. However, it has been demonstrated that a rotating plume without external forcing can be very stable over many rotational periods (Fong and Geyer 2002; Lentz and Helfrich 2002; Hetland 2005; Horner-Devine et al. 2006; Hetland 2017); this implies that the suppression can be due to the intrinsic inhibiting effects of the front. Baroclinic instabilities can enhance dispersion of water borne particles (Thyng and Hetland 2018), and decrease predictability in numerical simulations (Marta-Almeida et al. 2013). Better understanding the instability growth would improve our understanding on numerical predictability and be helpful for further investigating other submesoscale processes (e.g., symmetric instabilities and frontogenesis), material (e.g., nutrients and sediments) transport, mixing, and turbulence in the bottom boundary layers and coastal buoyancy-driven flows.
This paper explores nongeostrophic baroclinic instability theories adapted to the scenario with a sloping bathymetry. The two-layer model is adapted from Sakai (1989), and the continuously stratified model is an existing model adapted from Stone (1971) by Wenegrat et al. (2018). Both models are used to investigate the suppression of nongeostrophic baroclinic instabilities over sloping bathymetry, which is not revealed in the classical QG theories (Blumsack and Gierasch 1972; Mechoso 1980). In particular, this paper attempts to seek a nondimensional parameter for indicating the instability suppression and understand the underlying mechanism controlling the suppression in the buoyant flow regime.
2. Theory
a. Layered model of nongeostrophic baroclinic instability
Phillips (1954) transformed the continuously stratified fluid to a two-layer system and constructed the layered model of QG baroclinic instabilities. Sakai (1989) investigated the ageostrophic instabilities using an ageostrophic version of the Phillips model. We adapt the Sakai model to the scenario with sloping bottom and surface (hereinafter referred to as the adapted Sakai model). The schematics of the Sakai model and the adapted Sakai model are shown in Fig. 1. The adapted Sakai model is a rotating two-layer channel with the sloping topography and background flow in the thermal wind balance. Considering the time scale as f−1, the horizontal length scale as the Rossby deformation radius
subject to
where the subscript 1 denotes the variables in the upper layer, subscript 2 for the lower layer, u is the along-slope velocity, υ is the across-slope velocity, p is the pressure, ±Ymax are the across-slope boundaries,

Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Substituting a wave-like solution
b. Continuously stratified model of nongeostrophic baroclinic instability
Eady (1949) developed a continuously stratified QG framework to study baroclinic instabilities. Blumsack and Gierasch (1972) adapted the Eady model to the scenario with a sloping bottom (hereinafter referred to as the BG model). Mechoso (1980) extended the BG model to the scenario with a sloping surface. Stone (1966, 1970, 1971) extended the Eady model to the nongeostrophic limit and constructed the continuously stratified model of nongeostrophic baroclinic instabilities. We use the modified model (Wenegrat et al. 2018), which adapts the Stone model to the scenario with sloping bottom and surface (hereinafter referred to as the adapted Stone model). The schematic of the BG, Mechoso, and adapted Stone models are shown in Fig. 2. In the adapted Stone model, the coordinates are rotated to align with the sloping bottom. The background buoyancy has a constant vertical gradient N2 and a constant horizontal gradient M2, and the background flow is constrained by the thermal wind relation. Considering the time scale as f−1, the horizontal length scale as U/f, and the vertical length scale as H (see the primitive equations and scale analysis in appendix B), the dimensionless form of the equations governing the perturbations in the rotated coordinates is
subject to
where u is the along-slope velocity, υ is the across-slope velocity, and w is the slope-normal velocity, p is the pressure, b is the buoyancy, θ is the bottom slope angle,

Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Normalized growth rate (
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Noting that the surface is also assumed to be tilted, the adapted Stone model is intrinsically suitable for baroclinic instabilities in a tilted bottom boundary layer, but seems not directly applicable to the situation with a flat surface that is not parallel with the tilted bottom (e.g., coastal buoyancy-driven flow). One way is to adapt the Stone model to the scenario with a flat surface, but this presents two challenges for theoretical approaches: first, making the uniform depth assumption like Eq. (4) is theoretically invalid; second and subsequently, assumed solutions with a wave form in the across-slope direction become theoretically invalid. Consequently, instead of adapting the Stone model to the flat-surface scenario, we will address the feasibility of the adapted Stone model (Wenegrat et al. 2018) in the flat-surface cases, for example, baroclinic instabilities in a coastal buoyancy-driven flow over a continental shelf.
Assuming a wave-like solution
The normalized growth rate
With regard to the normalization in the adapted Sakai model, the wavenumber
3. Results
In this section, we will address the suppression of nongeostrophic baroclinic instabilities over sloping bathymetry from the temporal perspective—the growth rate of instabilities—and discuss the suppression mechanisms in the layered and continuously stratified models. The suppression problem will be addressed in the dimensionless space; the dimensionless parameter space is Ri–δ, where Ri represents the nongeostrophic effect and δ represents the effect of sloping bathymetry.
a. Suppression of instabilities in the layered model
Based on the adapted Sakai model, baroclinic instabilities start to be suppressed (meaning the growth rate of the instabilities is reduced), when the bottom slope increases (slope parameter δ increases from 0 to 0.5, see Fig. 1). Also, instabilities are found to be suppressed with decreasing bulk Richardson number Rib. The suppression of instabilities has opposite dependencies on Rib and δ. To demonstrate this, the maximum growth rate as a function of Rib and δ is calculated. The maximum normalized growth rate

(left) Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(left) Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(left) Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
A new parameter, the slope-relative Burger number Sr ≡ (M2/f2)(α + αp), is considered as the coefficient controlling the suppression of instabilities in the nongeostrophic case. Compared to the horizontal slope Burger number SH ≡ (M2/f2)α (Hetland 2017), the slope-relative Burger number Sr uses a bottom slope relative to the isopycnal slope αp. Thus, if the slopes are aligned, with the bottom parallel to the isopycnals, then Sr = 0. The slope-relative Burger number uses the layer thicknesses in the adapted Sakai model;
where δr is the slope-relative parameter,
with an interpretation similar to the slope-relative Burger number Sr. For clarity, hereafter 1 + δ will be written as δr.
The bulk form of the slope-relative Burger number is
The distribution of Sr in Rib–δ space (shown in the center panel of Fig. 3) exhibits a similar pattern as the maximum normalized growth rate
A strength of the Sakai model is to interpret instabilities in the framework of wave resonance based on the physical wave coordinates (Sakai 1989). For instance, Kelvin–Helmholtz instabilities are interpreted as the resonance of interacting gravity waves. In this study, we will only focus on the interactions between Rossby waves to interpret the suppression of baroclinic instabilities. The physical wave coordinates, then, only consist of the Rossby waves; since the Rossby wave coordinates are the Fourier basis (see appendix C), the projection onto the physical wave coordinates is equivalent to conducting the Fourier transform (Pedlosky 2013). The details of the Rossby wave resonance in the adapted Sakai model can be found in appendix C. The resonance rate
where

(left) Comparison of the normalized Rossby wave resonance rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(left) Comparison of the normalized Rossby wave resonance rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(left) Comparison of the normalized Rossby wave resonance rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
The underlying mechanism is that Sr influences the wave resonance by modifying the properties of the Rossby waves. According to appendix C, the normalized Doppler-shifted phase speed of the Rossby waves in both layers is
where

Normalized Doppler-shifted phase speed in the upper (red) and lower (blue) layers. Gray dashed lines show the linear trend (r2 = 0.678, p = 0.0). Schematics exhibit the wave resonance in the scenarios of low and high Sr.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

Normalized Doppler-shifted phase speed in the upper (red) and lower (blue) layers. Gray dashed lines show the linear trend (r2 = 0.678, p = 0.0). Schematics exhibit the wave resonance in the scenarios of low and high Sr.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Normalized Doppler-shifted phase speed in the upper (red) and lower (blue) layers. Gray dashed lines show the linear trend (r2 = 0.678, p = 0.0). Schematics exhibit the wave resonance in the scenarios of low and high Sr.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
b. Suppression of instabilities in the continuously stratified model
The suppression of instabilities through the reduction of instability growth rate is also found in the adapted Stone model (Wenegrat et al. 2018). Particularly, the suppression is significant when the regime shifts from QG to nongeostrophic. To demonstrate the suppression, the QG models, that is, the BG model (Blumsack and Gierasch 1972) and the Mechoso model (Mechoso 1980), are briefly reviewed here. In the BG model, the background state consists of a density field with constant vertical and horizontal buoyancy gradients and a velocity field constrained by the thermal wind relation. By assuming that the Rossby number Ro = U/(fL) ≪ 1 and Burger number Bu = (NH)/(fL) ~ 1 (leading to the Richardson number Ri = Bu2Ro−2 ≫ 1), the equations governing perturbations can be reduced to a single PDE about the QG potential vorticity. Then, the growth rate of the instabilities is analytically obtained by solving the associated eigenvalue problem; the normalized growth rate of the BG model is then
where
Figure 2 shows
Maximum growth rates predicted by the QG and adapted Stone models have different dependencies on the Richardson number Ri. The maximum normalized growth rate

Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Maximum normalized growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
A dimensionless number is sought for indicating the suppression of instabilities in the adapted Stone model. One requirement of the number is that it can reduce to the QG limit (Blumsack and Gierasch 1972; Mechoso 1980) and the nongeostrophic limit with flat bathymetry (Stone 1970). The following discussion reviews the dimensionless numbers
which is obtained based on the asymptotic analysis of Stone (1970). Figure 7 (right) shows the comparison between the numerical

(a) Normalized maximum growth rates,
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(a) Normalized maximum growth rates,
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(a) Normalized maximum growth rates,
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Noticing that Sr|Flat = Ri−1, the controlling number (1 + Ri−1)−1/2 might be the reduction of (1 + Sr)−1/2 in the flat-bottom case. The multiplication of

(a)
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(a)
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(a)
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
The fact that
and
where
which is the ratio of two different types of EPE sources, that is, the EKE conversion and APE conversion to EPE; larger δ indicates larger EKE conversion back to EPE and hence stronger suppression. On the other hand,
larger Sr indicates smaller EPE conversion to EKE and hence weaker instability growth.
It has been suggested by previous studies that the horizontal slope Burger number SH = δ/Ri is relevant to the instability suppression (Hetland 2017; Wenegrat et al. 2018). Wenegrat et al. (2018) demonstrate that VBFc/VBFn can be scaled as SH; larger SH indicates larger reduction of VBF and hence weaker instabilities. The difference of scaling is that SH scales the ratio between the VBF components rather than the components as Eqs. (17) and (18) do. Figure 8b shows SH in the Ri − δ space, and Fig. 8d shows the comparison between SH and
4. Discussion
a. Application to coastal buoyancy-driven flow
The adapted theories are directly applicable to baroclinic instabilities in a bottom boundary layer (Wenegrat et al. 2018), where both the bottom and surface can be assumed to be tilted and parallel. However, these theories may not seem as directly applicable to surface-intensified baroclinic instabilities that are usually associated with tilted bottoms but flat surfaces (e.g., instabilities in coastal fronts). In this section, we will address, by scale analysis, the applicability of the adapted Stone model to the baroclinic instabilities in coastal buoyancy-driven flow over sloping bathymetry.
The following scale analysis demonstrates that the adapted Stone model can be used as an approximation of the flat surface case so long as the slope Burger number
where S ≡ (N/f)α = δRi−1/2 is the slope Burger number. For the situations characterized by δ ~ O(10−1) and Ri ~ O(1) (which are the cases in this study), the slope Burger number S is O(10−1) so that the uniform fluid depth z = 1 will be a reasonable assumption. In other words, if S ~ O(10−1), motions with lateral displacements O(Rd) will span a depth range of H ± 0.1H, and hence the uniform fluid depth will be a first-order approximation.
With the uniform fluid depth approximation, the rigid-lid boundary condition at the flat surface in the rotated coordinates is
where SH ≡ αM2f−2= δRi−1 is the horizontal slope Burger number (Hetland 2017). For the situations characterized by δ ~ O(10−1) and Ri ~ O(1) (which are the cases in this study), the horizontal slope Burger number SH is O(10−1) so that the no-flow boundary condition
b. Numerical simulations of instabilities in coastal buoyancy-driven flow
A series of existing idealized numerical simulations, first analyzed by Hetland (2017), are used to examine the feasibility of the adapted Stone model. The idealized model domain is a 260 km (along-slope) × 128 km (across-slope) continental shelf with the uniform bathymetric slope α = 10−3 across all simulations. The depth increases from 5 m onshore to 133 m offshore. The model grid has 1-km uniform horizontal resolution and 30 layers in the vertical direction. The boundary conditions are periodic along-slope, open (with a sponge layer) offshore, and closed (no-slip) at the coast. The k–ε turbulence closure scheme is used to calculate the vertical mixing, and bottom friction is defined using a specified bottom roughness and a log-layer approximation. The model is unforced and run as an initial-value problem. The initial density field is a coastal buoyant front with a constant vertical stratification N2 over the whole shelf and a constant lateral buoyancy gradient M2 within the offshore distance of W = 50 km. The initial current field is configured in the thermal wind balance with the density field. The initial density and current fields of the base case (Ri = 2.0 and δ = 0.1) are shown in Fig. 9. Initial fields are varied among simulations to cover a range of situations of instability formation.

(top) Model domain and initial surface density and across-shore sections of initial (middle) density and (bottom) current (along the red dashed line in the top panel).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(top) Model domain and initial surface density and across-shore sections of initial (middle) density and (bottom) current (along the red dashed line in the top panel).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(top) Model domain and initial surface density and across-shore sections of initial (middle) density and (bottom) current (along the red dashed line in the top panel).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
The idealized simulations were configured in the parameter space of the Richardson number Ri and the slope parameter δ, and all the simulations used in this study are listed in Table 1. All the simulations were run with same stratification N2 and same bottom slope α, but with different lateral buoyancy gradients M2 and Coriolis parameters f that are determined by each combination of Ri and δ. Note that the simulations listed in Table 1 are part of the simulations conducted in Hetland (2017); simulations with the Richardson number Ri = 1 or Ri = 10 or the slope parameter δ > 0.5 or the horizontal slope Burger number SH > 0.2 were excluded. Ri = 1 is around the boundary between baroclinic instabilities and symmetric instabilities (Haine and Marshall 1998; Boccaletti et al. 2007), so the simulations with Ri = 1 are excluded to ensure only baroclinic instabilities can form. The simulations with Ri = 10 are excluded to ensure the nongeostrophic regime, and to minimize the influence of bottom friction in these simulations with long instability growth rates. The instability formations in the simulations with δ > 0.5 are excluded because they are generally weak and also strongly influenced by bottom friction. The simulations with SH > 0.2 are excluded to ensure
Simulations in the parameter space of Ri and δ. All simulations were run with N2 = 1.00 × 10−4 s−2 and α = 1.00 × 10−3. Other parameters are determined by varing Ri and δ. The slope Burger number is determined by S = δRi−1/2. The horizontal slope Burger number is determined by SH = δRi−1. The Coriolis parameter is determined by f = Nα/S. The horizontal buoyancy gradient is determined by M2 = NfRi−1/2. The terms



Development of instabilities of the base case (Ri = 2.0 and δ = 0.1) at the (left) surface and (right) bottom.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

Development of instabilities of the base case (Ri = 2.0 and δ = 0.1) at the (left) surface and (right) bottom.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Development of instabilities of the base case (Ri = 2.0 and δ = 0.1) at the (left) surface and (right) bottom.
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
The growth rate of instabilities is estimated in each simulation and then compared to the QG model and the adapted Stone model. Note that the spatial scale of the instabilities increases in the offshore direction (see Fig. 10) because the deformation radius, Rd = NH/f, increases with increasing depth offshore. We only focus on the growth of the largest instabilities at ~50 km offshore with the depth H ~ 50 m, because they are the most energetic and dispersive components (Thyng and Hetland 2017, 2018).
EKE is used to quantify the growth rates of instabilities. Given that the domain is periodic in the along-slope direction, the velocity field associated with the instability eddies is calculated by subtracting the along-slope background velocity from the original velocity field. So, EKE is dominated by the largest eddies. Then, the EKE can be determined by integrating the kinetic energy of the eddy flow field over the whole domain. Last, the EKE is normalized by the initial domain-integrated mean kinetic energy MKEInitial. Figure 11 (top) shows the normalized EKE, EKE/MKEInitial, of all the simulations listed in Table 1.

(top) Normalized EKE time series of the simulations of Table 1. (middle) Truncated normalized EKE time series. Each series is truncated at where the EKE reaches 50% of its maximum. Colors of the lines denote δ, and darker colors represent lower Ri. (bottom) Normalized EKE time series of the base case (Ri = 2.0, δ = 0.1) is compared to the theoretical estimates. The best exponential function fitting the base case has a growth rate of 1.73 day−1 (r2 = 0.996), and the theoretical estimates are 1.82 day−1 (nongeostrophic; adapted Stone model), 2.25 day−1 (nongeostrophic, flat bottom; Stone model), and 2.50 day−1 (QG; BG model).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(top) Normalized EKE time series of the simulations of Table 1. (middle) Truncated normalized EKE time series. Each series is truncated at where the EKE reaches 50% of its maximum. Colors of the lines denote δ, and darker colors represent lower Ri. (bottom) Normalized EKE time series of the base case (Ri = 2.0, δ = 0.1) is compared to the theoretical estimates. The best exponential function fitting the base case has a growth rate of 1.73 day−1 (r2 = 0.996), and the theoretical estimates are 1.82 day−1 (nongeostrophic; adapted Stone model), 2.25 day−1 (nongeostrophic, flat bottom; Stone model), and 2.50 day−1 (QG; BG model).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(top) Normalized EKE time series of the simulations of Table 1. (middle) Truncated normalized EKE time series. Each series is truncated at where the EKE reaches 50% of its maximum. Colors of the lines denote δ, and darker colors represent lower Ri. (bottom) Normalized EKE time series of the base case (Ri = 2.0, δ = 0.1) is compared to the theoretical estimates. The best exponential function fitting the base case has a growth rate of 1.73 day−1 (r2 = 0.996), and the theoretical estimates are 1.82 day−1 (nongeostrophic; adapted Stone model), 2.25 day−1 (nongeostrophic, flat bottom; Stone model), and 2.50 day−1 (QG; BG model).
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
The EKE in each case appears to increase exponentially from the start (Fig. 11, top), but eventually the rapid increase is retarded by friction and finite amplitude effects. To isolate our results from these frictional effects and compare our results more directly with the theories that do not consider the influence of friction, we truncate the EKE time series where it reaches half of the maximum, removing the later part that is potentially influenced by friction. The truncated time scale for each simulation is listed in Table 1, and the truncated EKE time series are shown in Fig. 11 (middle). We take the base case (Ri = 2.0, δ = 0.1) as an example to show the comparison between the simulated growth rate and the theoretical predictions. The truncated EKE time series of the base run is shown in Fig. 11 (bottom), and the best exponential function to fit it has a growth rate of 1.73 day−1 (r2 = 0.996). On the other hand, the maximum dimensional growth rate for the base case is estimated based on the adapted Stone model (Wenegrat et al. 2018), the flat-bottom Stone model (Stone 1971), and the QG model (Blumsack and Gierasch 1972); they are
A regressed estimate of simulated EKE growth rate,

(top) The regressed growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1

(top) The regressed growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
(top) The regressed growth rate
Citation: Journal of Physical Oceanography 50, 7; 10.1175/JPO-D-19-0145.1
Both the nongeostrophic effect and sloping bathymetry are important for indicating the instability growth. The overestimation of growth rates in QG theory is presumably due to the missing of the nongeostrophic suppressing effect. Applying the flat-bottom Stone theory misses the suppressing effect of the sloping bathymetry. From the perspective of energetics, assuming flat-bottom (δ = 0) will overestimate the total VBF (by underestimating the EKE sink VBFc and overestimating the EKE source VBFn) and hence underestimate the suppression. From the perspective of wave interactions, assuming the flat-bottom will facilitate/overestimate the wave interaction and hence underestimate the suppression as well.
Furthermore, the adapted Stone theory might overestimate the suppression in the coastal case by assuming a sloping surface, as the comparison between the BG and Mechoso models suggests (see Fig. 2). But the nonlinear simulations indicate that the adapted Stone theory is most accurate among these existing theories, which suggests that the overestimation is not prohibitively large. On the other hand, there is little improvement when moving from the QG theory to the flat-bottom Stone theory; the underestimation of the suppression by assuming a flat-bottom seems not negligibly small.
c. Linking Sr to potential vorticity
Above, the parameter Sr is linked to instability growth rate. Here we show how this parameter is related to normalized potential vorticity. The reference potential vorticity in this case is fN2, the case with no horizontal density gradients and thereby no associated balanced flow. In the adapted Stone model, the normalized Ertel potential vorticity (Thomas et al. 2016) can be written as
For the adapted Stone model, the potential vorticity is uniform throughout the domain, but in the adapted Sakai model, the sloping interface creates a gradient in potential vorticity. In this case, we consider the change in potential vorticity, ΔPV, across an inertial radius Li = U0/f
where βT = [(α + αp)f]/H0 is the topographic beta supplemented by the isopycnal slope. Thus, if we consider a parcel going from high to low potential vorticity, the normalized potential vorticity will be
5. Conclusions
Layered and continuously stratified models of nongeostrophic baroclinic instability over sloping topography are explored in the buoyant flow regime to study the suppression mechanism. The layered model (i.e., the adapted Sakai model) reveals a new parameter, the slope-relative Burger number Sr ≡ (M2/f2)(α + αp) {the bulk form is Sr = [U0/(H0f)](α + αp)}, which is a dimensionless parameter that controls the suppression of instabilities in the nongeostrophic limit. The continuously stratified model (i.e., the adapted Stone model) shows that, when the regime shifts from QG to nongeostrophic, the growth of instabilities is inhibited with increasing Sr.
In the adapted Sakai model, the instability growth rate linearly decreases with increasing Sr. The physical mechanism behind Sr is explored based on the wave resonance theory (Sakai 1989). In the physical wave coordinates consisting of the Rossby waves, baroclinic instabilities are interpreted as the Rossby wave resonance; supported by the adapted Sakai model, the maximum normalized growth rate of instabilities is found to be nearly 1:1 to the resonance rate of the Rossby waves. The slope-relative Burger number Sr modifies the Doppler-shifted phase speed of the Rossby waves, alters the wave resonance, and hence influences the growth of instabilities.
In the adapted Stone model, the instability growth rate linearly decreases with decreasing
The adapted models are intrinsically applicable to baroclinic instabilities in a bottom boundary layer but has one limitation when applying to coastal buoyancy-driven flow—the sloping-surface assumption. Supported by scale analysis, idealized numerical simulations of coastal buoyancy-driven flow are used to test the feasibility of the adapted Stone model in flat-surface situations. The comparison of the numerical results and theoretical predictions indicates that the limitation is not prohibitive if the slope Burger number
Acknowledgments
This work was funded by NSF (Grant OCE-1851470), Texas General Land Office (Contract 18-132-000-A673), and Gulf of Mexico Research Initiative through the CSOMIO consortium (Contract R01984). Lixin Qu was supported by a graduate research fund from Texas Sea Grant and a scholarship from China Scholarship Council. We thank Jacob Wenegrat, Leif Thomas, and two anonymous reviewers for very helpful suggestions and codes when preparing this manuscript.
APPENDIX A
Adapted Sakai Model
The following derivation follows Sakai (1989) but is modified to account for the presence of sloping bottom and surface. Considering a rotating two-layer channel with sloping bottom and top and currents in the thermal wind balance, the linearized equations of the perturbations are
subject to
where
Considering the time scale as 1/f, the horizontal length scale as the Rossby deformation radius
where
subject to
Assuming an ansatz of the form
subject to
APPENDIX B
Adapted Stone Model
The following derivation follows Stone (1966, 1970, 1971) but is modified to account for the presence of sloping bottom and surface. The coordinates are rotated to align with the sloping topography as Wenegrat et al. (2018), and the derivation is essentially equivalent to Wenegrat et al. (2018) but with a different coordinate orientation. Dimensionally, the equations describing a rotational fluid field with the Boussinesq approximation in the rotated coordinates are
subject to
where θ is the slope angle,
Considering the horizontal velocity scale as U, the time scale as f−1, the horizontal length scale as U/f, and the vertical length scale as H, the scaling relations about the variables in Eq. (B1) are, then,
Thus, Eqs. (B1) and (B2) have the dimensionless form
subject to
where Ri = N2H2U−2 = N2f2M−4 is the Richardson number, δ = αN2M−2 is the slope parameter, and ε = fHU−1 = f2M−2 is the nonhydrostatic parameter.
Considering a background current flowing only in the along-slope direction, constrained by the thermal wind relation, the background state can be described as
where M2/f is the thermal wind shear in the nonrotated coordinates and
where Sr = δrRi−1 is the slope-relative Burger number. Assuming u′, υ′, w′, b′, and p′ are the small perturbations from the background state, the equations governing the perturbed state can be linearized by neglecting the product of small terms as (dropping the primes for clarity and neglecting the viscosity and diffusion terms)
Assuming an ansatz of the form
subject to
APPENDIX C
Rossby Wave Interactions in the Adapted Sakai Model
The following derivation follows Sakai (1989) but is modified to account for the presence of sloping bottom and top. The details about the interaction theory and associated derivation can be found in section 4 and appendix A of Sakai (1989). We will only focus on the interactions between Rossby waves. In the adapted Sakai model, the physical wave coordinates consisting of the Rossby waves in the upper and lower layers are, then,
where
where An and Bn are the magnitudes in the physical wave coordinates,
where
which can be reduced to the flat-bottom case,
APPENDIX D
Energetics in the Rotated Coordinates
The following energetics closely follows Wenegrat et al. (2018) but with a different coordinate orientation. The coordinates are rotated to align with the sloping topography. The relation between the nonrotated and rotated coordinates is
where the tildes denote the nonrotated coordinates. The background buoyancy has a constant vertical gradient
So, the background buoyancy in the rotated coordinates could be set as
By linearizing the dimensional governing equations, Eq. (B1), we can get the dimensional equations describing the perturbations as below (dropping asterisks for clarity)
where the bars denote the background variables, the primes denote the perturbation variables, and
By multiplying b′/N2 to the buoyancy equation in Eq. (D3) and taking the Reynolds averaging, the eddy potential energy (EPE) equation can be obtained as
where HBFc is the horizontal buoyancy flux (HBF) contributed by cross-slope motion, HBFn is the HBF contributed by slope-normal motion, VBFc is the vertical buoyancy flux (VBF) contributed by cross-slope motion, VBFn is the VBF contributed by slope-normal motion, DPE is the dissipation of EPE, and RPE is the redistribution of EPE. Baroclinic instabilities get energy from the available potential energy of the background front. The energy transfer from mean available potential energy to EPE is through the HBF terms, which requires
By multiplying
where SP is the shear production, DKE is the dissipation of EKE, RKE is the redistribution of EKE, and
REFERENCES
Blumsack, S. L., and P. Gierasch, 1972: Mars: The effects of topography on baroclinic instability. J. Atmos. Sci., 29, 1081–1089, https://doi.org/10.1175/1520-0469(1972)029<1081:MTEOTO>2.0.CO;2.
Boccaletti, G., R. Ferrari, and B. Fox-Kemper, 2007: Mixed layer instabilities and restratification. J. Phys. Oceanogr., 37, 2228–2250, https://doi.org/10.1175/JPO3101.1.
Brannigan, L., D. P. Marshall, A. C. Naveira Garabato, A. G. Nurser, and J. Kaiser, 2017: Submesoscale instabilities in mesoscale eddies. J. Phys. Oceanogr., 47, 3061–3085, https://doi.org/10.1175/JPO-D-16-0178.1.
Burns, K. J., G. M. Vasil, J. S. Oishi, D. Lecoanet, and B. Brown, 2016: Dedalus: Flexible framework for spectrally solving differential equations. Astrophysics Source Code Library, http://www.ascl.net/1603.015.
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.
Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1, 33–52, https://doi.org/10.3402/tellusa.v1i3.8507.
Fong, D. A., and W. R. Geyer, 2002: The alongshore transport of freshwater in a surface-trapped river plume. J. Phys. Oceanogr., 32, 957–972, https://doi.org/10.1175/1520-0485(2002)032<0957:TATOFI>2.0.CO;2.
Fox-Kemper, B., R. Ferrari, and R. Hallberg, 2008: Parameterization of mixed layer eddies. Part I: Theory and diagnosis. J. Phys. Oceanogr., 38, 1145–1165, https://doi.org/10.1175/2007JPO3792.1.
Haine, T. W., and J. Marshall, 1998: Gravitational, symmetric, and baroclinic instability of the ocean mixed layer. J. Phys. Oceanogr., 28, 634–658, https://doi.org/10.1175/1520-0485(1998)028<0634:GSABIO>2.0.CO;2.
Hetland, R. D., 2005: Relating river plume structure to vertical mixing. J. Phys. Oceanogr., 35, 1667–1688, https://doi.org/10.1175/JPO2774.1.
Hetland, R. D., 2017: Suppression of baroclinic instabilities in buoyancy-driven flow over sloping bathymetry. J. Phys. Oceanogr., 47, 49–68, https://doi.org/10.1175/JPO-D-15-0240.1.
Horner-Devine, A. R., D. A. Fong, S. G. Monismith, and T. Maxworthy, 2006: Laboratory experiments simulating a coastal river inflow. J. Fluid Mech., 555, 203–232, https://doi.org/10.1017/S0022112006008937.
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.
Lentz, S. J., and K. R. Helfrich, 2002: Buoyant gravity currents along a sloping bottom in a rotating fluid. J. Fluid Mech., 464, 251–278, https://doi.org/10.1017/S0022112002008868.
Marta-Almeida, M., R. D. Hetland, and X. Zhang, 2013: Evaluation of model nesting performance on the Texas-Louisiana continental shelf. J. Geophys. Res. Oceans, 118, 2476–2491, https://doi.org/10.1002/jgrc.20163.
Mechoso, C. R., 1980: Baroclinic instability of flows along sloping boundaries. J. Atmos. Sci., 37, 1393–1399, https://doi.org/10.1175/1520-0469(1980)037<1393:BIOFAS>2.0.CO;2.
Pedlosky, J., 2013: Geophysical Fluid Dynamics. Springer Science & Business Media, 710 pp.
Pedlosky, J., 2016: Baroclinic instability over topography: Unstable at any wave number. J. Mar. Res., 74, 1–19, https://doi.org/10.1357/002224016818377595.
Phillips, N. A., 1954: Energy transformations and meridional circulations associated with simple baroclinic waves in a two-level, quasi-geostrophic model. Tellus, 6, 274–286, https://doi.org/10.3402/tellusa.v6i3.8734.
Sakai, S., 1989: Rossby-Kelvin instability: A new type of ageostrophic instability caused by a resonance between rossby waves and gravity waves. J. Fluid Mech., 202, 149–176, https://doi.org/10.1017/S0022112089001138.
Stone, P. H., 1966: On non-geostrophic baroclinic stability. J. Atmos. Sci., 23, 390–400, https://doi.org/10.1175/1520-0469(1966)023<0390:ONGBS>2.0.CO;2.
Stone, P. H., 1970: On non-geostrophic baroclinic stability: Part II. J. Atmos. Sci., 27, 721–726, https://doi.org/10.1175/1520-0469(1970)027<0721:ONGBSP>2.0.CO;2.
Stone, P. H., 1971: Baroclinic stability under non-hydrostatic conditions. J. Fluid Mech., 45, 659–671, https://doi.org/10.1017/S0022112071000260.
Thomas, L. N., J. R. Taylor, E. A. D’Asaro, C. M. Lee, J. M. Klymak, and A. Shcherbina, 2016: Symmetric instability, inertial oscillations, and turbulence at the gulf stream front. J. Phys. Oceanogr., 46, 197–217, https://doi.org/10.1175/JPO-D-15-0008.1.
Thyng, K. M., and R. D. Hetland, 2017: Texas and Louisiana coastal vulnerability and shelf connectivity. Mar. Pollut. Bull., 116, 226–233, https://doi.org/10.1016/j.marpolbul.2016.12.074.
Thyng, K. M., and R. D. Hetland, 2018: Seasonal and interannual cross-shelf transport over the Texas and Louisiana continental shelf. Cont. Shelf Res., 160, 23–35, https://doi.org/10.1016/j.csr.2018.03.006.
Wenegrat, J. O., J. Callies, and L. N. Thomas, 2018: Submesoscale baroclinic instability in the bottom boundary layer. J. Phys. Oceanogr., 48, 2571–2592, https://doi.org/10.1175/JPO-D-17-0264.1.