1. Introduction
In recent decades, warm ocean circulation within ice shelf cavities has led to the accelerated melting of glaciers in West Antarctica and northern Greenland (Schodlok et al. 2012; Mayer et al. 2000). This circulation contributes to glacial melting through the inward transport of warm, salty water masses and amplification of the ice–ocean heat flux through greater flow velocities and temperature gradients (Holland et al. 2008). Recent observations and modeling have also shown that as glaciers retreat past bathymetric maxima, the bathymetry has a leading-order effect on the glacial melt rates as subglacial cavities are shaped by oceanic circulations on the undersides of floating glaciers (Gudmundsson et al. 2012). This rapid retreat, postulated to be caused by the retrograde bathymetric slopes (increasing elevation in the direction of ice flow), has been observed using satellite radar interferometry to varying degrees in West Antarctic glaciers such as the Pine Island, Thwaites, Smith, and Kohler glaciers (Rignot et al. 2014). Autosub cavity-transect measurements in the Pine Island Glacier (PIG) revealed a cavity that hosts the most prominent bathymetric sill yet observed and exhibits the fastest melt rate in the region (Jenkins et al. 2010; Kimura et al. 2016). Floating glaciers in northern Greenland such as the 79 North Glacier (whose sill is external to the cavity) and Petermann Glacier have similar geometries and oceanic circulations (Cai et al. 2017; Schaffer 2017).
Previous studies have suggested that grounding lines retreating on a descending slope will ultimately lead to marine ice sheet instability, a dynamically driven, runaway retreat that can result in complete discharge of the ice stream (Schoof 2007; Gudmundsson 2013; Joughin et al. 2014). The bathymetric influence of many of the glaciers in the Amundsen Sea in West Antarctica, with the PIG as a prime example, are critical in understanding why this region is the fastest melting sector of Antarctica. Here, the melting occurs primarily at the ice–ocean interface and is predominantly forced by warm ocean circulation, as opposed to calving processes or surface melting leading to subglacial discharge (Mouginot et al. 2014; Konrad et al. 2017). The sill under the PIG is speculated to have a controlling effect on melt rates due to its modulation of intruding warm, salty Circumpolar Deep Water (CDW) into Pine Island Bay (De Rydt et al. 2014). Data and observations inside the cavity have been sparse, but diabatic processes deep inside such an ice shelf cavity lead to persistent differences between the water mass properties in the interior compared to the open ocean (Jacobs et al. 2011; Dutrieux et al. 2014).
Previous studies have investigated the dynamics using simplified models with idealized ice shelf geometries without sills (Little et al. 2008) and with sills (De Rydt et al. 2014; De Rydt and Gudmundsson 2016), as well as using more comprehensive regional ocean models with realistic bathymetry (Schodlok et al. 2012; Nakayama et al. 2014; St-Laurent et al. 2015; Seroussi et al. 2017). In these studies, the bathymetric sill was consistently found to act as a topographic barrier to the inflow of warm, salty CDW, provided that the CDW layer was sufficiently thin. These models all predict enhanced friction at the ice–ocean interface as a result of the sill, which leads to greater melt rates.
The problem of pressure-driven flow across a bathymetric obstruction has been widely studied in other oceanographic contexts, such as hydraulically controlled flows (Whitehead et al. 1974; Gill 1977). However, this characterization is also appropriate for glaciers like the PIG as bathymetrically modulated exchange flows. The establishment of hydraulic control is a potential explanation for the sill’s apparent role as an obstruction to CDW inflow mentioned in previous studies (De Rydt et al. 2014; Dutrieux et al. 2014) and is discussed in section 7.
Under an ice shelf, a simple model for the leading-order dynamics is an exchange flow forced by negative buoyancy on the open-ocean side and a positive buoyancy in the far interior. This represents the common situation in West Antarctic and northern Greenland glaciers where a denser bottom layer is thicker on the open-ocean side than in the cavity, forcing an inflow at depth via a pressure head (Dutrieux et al. 2014). The oceanic buoyancy is then increased by freshening of the water mass through a transformation at the ice–ocean boundary occurring due to subglacial melting at the ice–ocean boundary in the far interior of the cavity, which establishes an overall isopycnal tilt along the length of the cavity.
Previous studies have placed less emphasis on the dynamical mechanisms via which the sill constrains the circulation. As a result, relatively little is understood about the flow regimes that manifest in cavity circulation and the important physical parameters that define these regimes and control the cross-sill transport. Using a high-resolution, minimal model, we simulate the circulation patterns over varying flow parameters to provide a qualitative and quantitative understanding of the transport and forms of variability that can be expected in real ice shelf cavities.
The outline of the paper is as follows. In section 2, we discuss the setup of a simplified two-layer dynamical model and the details of the numerical methods used to study this problem. We also discuss the considerations for a robust posing of the idealized problem. In section 3, we discuss the reasoning for partitioning the parameter space into three regimes based on the nondimensionalized sill height and friction velocity. In section 4, we present solutions in the high-friction (HF) regime and explain this theoretically as a friction-dominated Stommel balance regime with equations for the boundary layer and a prediction for the flow structure. In section 5, we present solutions in the low-friction, low-sill (LFLS) regime and discuss the emergence of gyres and eddies using potential vorticity (PV) and energy budgets. In section 6, we present solutions in the low-friction, high-sill (LFHS) regime and discuss the phenomena of shocks, which are sharp interface gradients that appear due to wave propagation in critical flow, and layer-grounding, which occurs when the bottom layer thickness reaches zero. In section 7, we derive theoretical estimates of the cross-sill transport based on rotating hydraulic control theory with uniform PV, discuss their limitations, and compare with numerical results. We classify and discuss the analytical and numerical boundaries between the regimes as a function of friction and sill height. In section 8, we summarize our findings and discuss the implications and caveats of the study. In section 9, we conclude and discuss future research directions.
2. Idealized cavity flow
a. A two-layer model




















The layer thicknesses at the north and south boundaries are linearly restored toward prescribed reference thicknesses, providing a simplistic representation of processes that occur outside of the cavity and the interior. The water mass transformation near ice shelves can be significant wherever ocean waters with temperatures above the local freezing point are in contact with the ice shelf. Melting is typically enhanced near grounding lines, which for Antarctic ice shelves are partly due to the greater depths for which the pressure-dependent freezing point is reduced to lower temperatures (Joughin et al. 2012). Also, in some warm water cavities such as the PIG, which has relatively warm (i.e., above surface freezing temperature) intrusions, the location at which the ice is in contact with the warmest water is primarily concentrated at the grounding line (Dutrieux et al. 2014). Our simplification is valid only for glaciers where the sill maximum is not located near the primary source of water mass transformation. For more general lock-exchange applications, this water mass transformation, which effectively fixes the stratification, represents any external processes that do not occur near the topographic sill as long as it leads to a relatively steady across-sill pressure head.
We model the two-layer channel flow problem purely dynamically and do not include interior diabatic mixing or thermal fluxes between layers or to the ice shelf and bottom boundary, except for a simplified representation of water mass transformation at the northern and southern boundaries. This means that although we are motivated by a flow under an ice shelf cavity, our study is not specific to ice–ocean interactions and the results also apply to channel flows in general. Because of this simplified dynamical framework, we do not make any predictions for the PIG melt rate, but a total transport can be used to constrain melt rate estimates via the water mass transformation. Instead, we focus on the circulation patterns that emerge and the bathymetric and geometric constraints on these patterns and the resulting transport. Also, the depth distribution of the water mass transformation/diabatic forcing is an important topic, but not addressed in the present study, and would be better addressed with more complete vertical resolution than in the two-layer model here.















The cavity has dimensions
















Using representative values for the PIG (Jacobs et al. 2011), the Coriolis parameter is
b. Numerical methods
To run simulations of the channel flow problem, we use the Back of Envelope Ocean Model (BEOM), a publicly available FORTRAN code written by Pierre St-Laurent (St-Laurent 2018). The numerical scheme is similar to the Hallberg Isopycnal Model (Hallberg and Rhines 1996) but offers a special treatment of layer-grounding (isopycnal outcropping when layer thicknesses vanish) using a Salmon layer (Salmon 2002). The Salmon layer introduces an artificial term added to the Montgomery potential that prevents numerical instability due to layer-grounding by raising the potential energy of the layer to infinity as its thickness approaches zero. PV is conserved in the continuous equations with the Salmon layer present.
We modify this code to include a rigid lid pressure solver, variable rigid lid elevation, friction against the rigid lid, and biharmonic viscosity. We assume equal top and bottom friction velocities in this study. The model uses the generalized forward–backward scheme (Shchepetkin and McWilliams 2005), which has been modified for compatibility with the rigid lid implementation and is discussed in appendix B. The boundary conditions are free-slip and closed on all four boundaries (no normal flow), with thickness nudging at the north and south boundaries, as mentioned in section 2a. The model uses energy-conserving finite differences (Sadourny 1975) for momentum and third-order upwinding (e.g., Shchepetkin and McWilliams 1998) for thickness advection on an Arakawa C grid of uniform resolution
All time-averaged results in the following discussions are derived from 100-day averages at the end of each simulation, which is approximately one residence time scale, defined by the cavity volume divided by the exchange rate (
c. Geostrophic transport
With the boundary condition posing in section 2a, there is a fixed isopycnal tilt along the channel if a strong enough nudging is used. Assuming the end points of the stratification are essentially fixed by the nudging, we can make an estimate for the zonal geostrophic transport. Here zonal (cross channel) geostrophic transport is used as a reference scale for the meridional (along channel) geostrophic transport; generally all of the flow crosses the channel from west to east as it flows from north to south due to topographic steering by the sill, so the zonal geostrophic transport and meridional geostrophic transport are approximately equal. This does not mean the cross-channel and along-channel pressure gradients need always be the same, but the reference zonal transport is found to predict the net cross-sill exchange well in most of the experiments discussed in sections 4–7.














3. Regime partitioning
For both our analytical and numerical study, we primarily examine the dynamics in a parameter regime relevant to the PIG, but our results also apply to the general class of nearly geostrophic, wide-channel flows, in which the channel width is much larger than the deformation radius
We show a reference case in Fig. 1, which illustrates the geometry of the model and the interface elevation η for a case with sill height

A representative geometry of a low-friction, high-sill case, with a snapshot of the interface elevation η for case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

A representative geometry of a low-friction, high-sill case, with a snapshot of the interface elevation η for case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
A representative geometry of a low-friction, high-sill case, with a snapshot of the interface elevation η for case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1


















Summary of key parameters for the numerical simulations and their corresponding estimates for the PIG. The sill height, friction velocity, and pressure head are varied independently in this study. The simulations will be referenced in the format


The chosen dimensionless
To understand variations of the parameters controlling the circulation, we study the effects of varying nondimensionalized friction, sill height, and pressure head from Eqs. (10a)–(10c). We plot the mean and root-mean-square deviation (RMSD) of the transport as a function of these three variables in Fig. 2 in which each parameter is varied separately, relative to a reference state defined by

Mean and RMSD of the nondimensionalized transport
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Mean and RMSD of the nondimensionalized transport
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Mean and RMSD of the nondimensionalized transport
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
In these results, variations in friction cause a 100% change to the variability over the studied range, while sill height causes transport to drop to 25% of QG prediction at
In the next three sections, we discuss each of these regimes and describe the phenomena that emerge, both qualitatively and quantitatively.
4. High-friction regime [
]

a. PV balance







Numerical solution for the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Numerical solution for the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Numerical solution for the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1









Terms in the thickness-weighted PV equation [Eq. (13)] for the bottom layer of the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Terms in the thickness-weighted PV equation [Eq. (13)] for the bottom layer of the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Terms in the thickness-weighted PV equation [Eq. (13)] for the bottom layer of the high-friction case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
b. Stommel boundary layer



















c. Scaling for the importance of bottom friction
The transition from high-friction to low-friction cases is determined by friction being





































5. Low-friction, low-sill regime (
, 
)


a. Transition to low friction: Gyres and eddies
In the HF cases, the numerical and analytical solutions predict a meridional antisymmetry about the domain center (due to a symmetric sill with equal magnitude and opposite sign in

Snapshot of interface elevation η at day 1000 for varying friction (low friction, low sill to high friction) cases (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Snapshot of interface elevation η at day 1000 for varying friction (low friction, low sill to high friction) cases (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Snapshot of interface elevation η at day 1000 for varying friction (low friction, low sill to high friction) cases (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Figure 6 shows the bottom layer relative vorticity

Snapshot of bottom layer vorticity
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Snapshot of bottom layer vorticity
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Snapshot of bottom layer vorticity
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
We show the time-averaged bottom layer PV for various values of

Bottom layer potential vorticity [defined in Eq. (11b)] for the same cases shown in Fig. 5—
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Bottom layer potential vorticity [defined in Eq. (11b)] for the same cases shown in Fig. 5—
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Bottom layer potential vorticity [defined in Eq. (11b)] for the same cases shown in Fig. 5—
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1







We plot the terms of the PV balance in Fig. 8 for the LFLS case

Terms in the thickness-weighted PV balance in Eq. (25) for LFLS case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Terms in the thickness-weighted PV balance in Eq. (25) for LFLS case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Terms in the thickness-weighted PV balance in Eq. (25) for LFLS case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
b. Eddy energetics
















Zoom of conversion terms in the eddy generation region comparing low-friction, low-sill and low-friction, high-sill cases
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Zoom of conversion terms in the eddy generation region comparing low-friction, low-sill and low-friction, high-sill cases
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Zoom of conversion terms in the eddy generation region comparing low-friction, low-sill and low-friction, high-sill cases
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
The LFLS panels in Fig. 9 show that both the barotropic and baroclinic eddy energy sources are concentrated close to the western boundary sill maximum. The horizontally integrated energy production is approximately 40% barotropic and 60% baroclinic, with a small region of intense negative barotropic conversion just before a localized peak in baroclinic conversion. Because of the sill, the bottom layer domain-integrated dissipation [calculated from the nonconservative terms in the thickness-weighted momentum equations from Aiki et al. (2016)] accounts for ⅔ of the total energy dissipated, while the top layer accounts for the remaining ⅓, since the bottom layer is more strongly eddying and the EKE dissipation by friction is roughly proportional to the EKE itself.
6. Low-friction, high-sill regime [
, 
]


a. Varying sill height
Increasing the sill height causes the boundary current velocities to increase, which strengthens the eddy intensity, as shown in Fig. 6. However, the bottom layer thickness decreases, so the vertically integrated transport and energy conversions generally decrease. The dynamics of the separation region of the western boundary current near the sill maximum change in structure and as seen in Fig. 2, the transport variability increases for intermediate sill height, and then decreases for the tallest sills. Figure 6 shows the bottom layer relative vorticity
Increasing the sill height also decreases the water column thicknesses at the sill maximum substantially, and most of this decrease occurs in the bottom layer. Therefore, the minimum width of boundary currents, which scales as
b. Shock formation
Shocks (or hydraulic jumps) are sharp interface gradients that arise when wave steepening due to nonlinear propagation occurs faster than any counteracting wave dispersion or dissipative processes (Helfrich et al. 1999). We observe the existence of standing shocks in our LFHS cases (defined here as the isopycnal steepness in a localized region of width
For the highest sills

Zoom view of
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Zoom view of
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Zoom view of
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1






In Fig. 11, we plot G in the vicinity of the sill maximum for three different values of

Zoom view of composite Froude number G for varying friction and fixed sill height
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Zoom view of composite Froude number G for varying friction and fixed sill height
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Zoom view of composite Froude number G for varying friction and fixed sill height
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
In our LFHS cases, we find that the meridional velocities are much larger than the zonal velocities near the shock region, similar to the dynamics in the western boundary layer before reaching the shock. The velocity field near the shock is approximately semigeostrophic (not shown); that is, meridional velocities across the shock are in near-geostrophic balance (with small frictional and Salmon layer contributions), while the zonal momentum balance has a leading-order advective contribution.
Even though the boundary current mainly flows around the separated region for these wide channel cases, we find that the bathymetric form drag of the shock acts to reduce the overall transport. The influence of form drag on the transport is further discussed in section 7a.
The small transport contribution within the grounded region is due to the Salmon layer, discussed in section 2b. The Salmon layer term in the region where the bottom layer
The positive vorticity region south of the sill coincides with the location of separation of the western boundary current and elevated barotropic conversion shown in Fig. 9. The location of the shock oscillates meridionally along the western boundary with period of days and distance
Since a steady-state viewpoint of the existence of Kelvin wave shocks appearing under critical conditions necessarily involves hydraulic control, we discuss the theory involved and how this influences the cross-sill exchange in the next section.
7. Constraining cross-sill exchange
In Fig. 2, transport is shown to be insensitive to friction, while sill height significantly reduces transport for
Additionally, the central gyre-like recirculation strength (which varies proportionally with the western boundary current) increases for low-friction cases. The recirculation strength reaches 3 times the total transport for the lowest friction case (shown in Fig. 7). There are also weaker gyre-like recirculations that develop in the northern and southern regions of the domain in response to the central gyre and connect it to the nudged regions in Fig. 7. The existence and strength of the gyre depend on a cross-isobath PV flux (unknown a priori), which is balanced by frictional vorticity destruction. However, the overall transport is approximately determined by only the forcing strength and the sill height, which leads to the difference in recirculation and overall transport.
The time scale needed to establish the total transport is on the order of tens of days (a few recirculation time scales, which is defined by the approximate time the current takes to cross from the northern to the southern boundary), while the western boundary current/recirculation strength takes hundreds of days to achieve equilibrium (a few residence time scales; not shown). In ice shelf cavities, the boundary layer transport influences the gyre-like recirculation strength, while the overall transport exerts a more direct control on the heat flux brought to the ice sheet (ignoring mixing processes, which may be important).
a. Hydraulic control theory
In this section, we test three different theoretical constraints for the cross-sill transport against the numerically diagnosed transports. These theoretical predictions rely on a steady-state analysis in the limit of no friction, often used for hydraulically controlled flows. Our aim is to predict the LFHS cross-sill exchange, which is important since the net cross-sill exchange most directly relates to basal melt.













We also derive a more comprehensive critical transport estimate for a two-layer rotating uniform PV assumption within a fixed width
For the highest sills, we can also predict the transport required for layer-grounding to occur, by considering a one-layer uniform-PV boundary layer of dynamic width
We compare the geostrophic, rotating one-layer zero PV, rotating two-layer uniform PV, and dynamic width transport estimates to the numerical results in Fig. 12. This shows that even the rotating one-layer zero PV version of the theory is reasonably accurate, particularly in the LFHS regime, since the Froude number in the bottom layer is generally much larger than the top-layer Froude number for higher sills

Theoretically predicted transport based on quasigeostrophic [Eq. (9)], critical, and grounded conditions, compared with time-averaged transport from our numerical simulations [with a gray dotted line excluding the IFD term in Eq. (6)] with weakest friction r = 1 × 10−5 m s−1, which corresponds to
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Theoretically predicted transport based on quasigeostrophic [Eq. (9)], critical, and grounded conditions, compared with time-averaged transport from our numerical simulations [with a gray dotted line excluding the IFD term in Eq. (6)] with weakest friction r = 1 × 10−5 m s−1, which corresponds to
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Theoretically predicted transport based on quasigeostrophic [Eq. (9)], critical, and grounded conditions, compared with time-averaged transport from our numerical simulations [with a gray dotted line excluding the IFD term in Eq. (6)] with weakest friction r = 1 × 10−5 m s−1, which corresponds to
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
For low sills, the transport matches the quasigeostrophic scaling [Eq. (9)] and decreases when critical conditions are reached. The quasigeostrophic scaling regime applies when the sill does not strongly affect the circulation strength and is also an upper bound that slightly overestimates the transport. One factor that accounts for this overestimation is the boundary nudging, which establishes a north–south interface gradient slightly smaller than the prescribed
For intermediate sill heights where
An alternative way of understanding the transport reduction is that for LFHS cases, the flow becomes asymmetric across the sill and produces appreciable form drag terms, which were previously assumed to be small in Eq. (6). The bathymetric and interfacial form drags contribute to the balance of the along-channel pressure force due to the imposed horizontal stratification, and thereby reduce the zonal (cross-sill) transport. The interface has approximate rotational symmetry in the HF cases, but does not in the LFLS cases. In both these cases, the interface does not exhibit large vertical excursions unless a shock forms, which is necessary for an appreciable form drag.
We calculate the relative contribution of transport reduction by IFD and BFD. The gray, dotted line in Fig. 12 shows
b. A regime diagram for cross-sill circulation
Based on the dynamical regimes discussed in the preceding sections, we now characterize the regimes of cross-sill flow over the entire parameter space of nondimensionalized friction and sill height parameter space in Fig. 13, which includes representative snapshots of the instantaneous bottom layer thickness. This diagram summarizes the three regimes that have been discussed in sections 4–6 and the analytical predictions and numerical results that define the regime boundaries.

Diagram of the three regimes discussed in sections 4–6 over nondimensionalized friction and sill height with insets showing illustrative snapshots of the interface elevation η. The numerical runs were categorized into these three regimes with different markers, as shown in the legend. The critical regime is classified as having a maximum composite Froude number
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Diagram of the three regimes discussed in sections 4–6 over nondimensionalized friction and sill height with insets showing illustrative snapshots of the interface elevation η. The numerical runs were categorized into these three regimes with different markers, as shown in the legend. The critical regime is classified as having a maximum composite Froude number
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Diagram of the three regimes discussed in sections 4–6 over nondimensionalized friction and sill height with insets showing illustrative snapshots of the interface elevation η. The numerical runs were categorized into these three regimes with different markers, as shown in the legend. The critical regime is classified as having a maximum composite Froude number
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
We empirically classify each individual simulation, which is represented as a point on the regime diagram, as critical if the maximum composite Froude number satisfies max(
Similarly, we classify individual simulations as eddying or friction-dominated based on the thickness-weighted PV Eq. (25). When the ratio of the friction to eddy terms spatially integrated and time averaged over the last 100 days of the run exceeds 1 {
We note that there is no explicit prediction for the boundary between the critical regime and the friction-dominated regime since friction does not have a strong effect on the transport, as discussed in section 3. However, we have extended the
8. Summary and discussion
In this study, we model the dynamics of ocean circulation in sill-impeded subshelf cavities, motivated by the fast-melting PIG (Jenkins et al. 2010; Dutrieux et al. 2014). To investigate the dynamics, we simplified the problem to a two-layer diabatically forced circulation with a simple channel geometry overlying a bathymetric sill. Our numerical solutions exhibit a variety of phenomena, including Stommel boundary layers, gyre-like circulations, eddies, hydraulic shocks, and layer-grounding.
This study demonstrates that friction and sill height have important controlling effects on the flow structure and variability of the flow in a simple buoyancy-forced flow. Increasing the friction does not strongly influence the transport, but decreases the variability. The transport under ice shelves is approximately geostrophic and generally controlled by pressure head and local water mass transformations due to the interaction with the ice base, except when the sill height begins to impede the flow. In the case of the PIG, this occurs for pycnocline depths
The theory for the HF regime provides an analytical solution for the flow profile and western boundary current over the sill, which narrows with larger
In the LFLS regime, barotropic/baroclinic eddies are abundant in the cavity, and reshape the flow into gyre-like circulations with strong inertial western boundary currents. We showed in section 5a that the inertial boundary layer and eddy PV fluxes, rather than friction, facilitate flow across the sill-imposed PV gradient. The central gyre-like recirculation does not reach the northern and southern nudged boundaries, but may still be important for diabatic mixing processes in the interior especially over the sill maximum (which may indirectly contribute to the heat transfer into and out of the domain and locally increase the heat exchange at the ice base).
In the LFHS regime discussed in section 6, standing shocks emerge in the numerical solutions when the flow reaches criticality, and either an increase in sill height or a shock amplitude increase due to low friction can lead to layer-grounding. The composite Froude number provides justification of the shock as an arrested Kelvin wave for the lowest friction cases.
The HF and LFLS cases generally exhibit a constant transport predicted by QG, while in the LFHS regime, the transport is sill-constrained and is reduced in accordance with hydraulic control theory. A regime diagram (Fig. 13) summarizes the HF, LFLS, and LFHS regimes. The dynamics of various glaciers in Antarctica and Greenland may be studied with their position in the regime diagram in mind. Specifically, the latest measurements in the PIG suggest that the sill-controlled circulation is in the LFHS critical regime, based on microstructure estimates that suggest
For ice shelf cavities, an additional way for flow to cross mean PV contours in the absence of eddies and bottom friction is lateral friction against the sidewalls (e.g., Little et al. 2008), but typically the depth tapers to zero at lateral boundaries in an ice shelf cavity (Kimura et al. 2016) and the top/bottom friction becomes dominant instead. The importance of lateral friction for each scenario can be formally calculated by finding its associated boundary length scale and comparing with
Given the transport and
Although this study is primarily motivated by the hydraulically controlled sill under an ice shelf cavity in the PIG, the regimes observed in Fig. 13 can also be generalized to many more sill-influenced exchange flows. Specifically, the top-layer friction does not strongly influence the regime diagram, since this layer is weakly modified by the sill and exhibits much weaker velocities (the composite Froude number is dominated by the bottom layer unless the top layer is much shallower). Therefore, these results can be directly applicable to scenarios with a free surface in open-ocean, wide-sill overflows with similar stratifications such as the Denmark Strait and the Faroe–Bank Channel (Pratt and Whitehead 2007), among many others.
Given the highly idealized nature of the study, there are numerous caveats associated with our results. In addition to geometric simplifications and quadratic friction (discussed in appendix A), there is no direct representation of ice–ocean thermodynamic interactions, no tidal flows, and no tidal mixing in the cavity. Also, the numerical model implements an advection and spatial discretization scheme that is not optimized for the study of shocks. However, with high horizontal and temporal resolution, and an observed steadiness in shock location, we are able to resolve these features with improved accuracy. Layer-grounding is treated in a special way in our simulations using Salmon layers, which prevent instabilities involving layer-grounding.
9. Conclusions and future directions
Our results corroborate well with previous modeling studies with idealized geometry (De Rydt et al. 2014; De Rydt and Gudmundsson 2016), and our idealized posing and high-resolution sweep over the key dimensionless parameters, allowing an exhaustive exploration of the dynamics. Based on this study, there are direct implications for future observations under ice shelves. These predictions may be used to guide measurements of the flow properties in ice shelf cavities and look for western boundary currents and shocks that may cause large diabatic fluxes, observable in western boundary currents downstream of the sill maximum. Such a region is likely to be a location of elevated mixing due to the sharp isopycnal tilt that forms when the bottom layer thins as it flows over the sill. The thinness of this layer and measured transport should have an influence on the magnitude of the shock amplitude and therefore, the turbulent mixing of subglacial water masses. Enhanced mixing near the sill is also likely to cause dynamical feedbacks; that is, when an initial pressure head inducing a circulation bringing heat toward the ice shelf causes increased water mass transformation, this further increases the pressure head. Measurements indicate that melting is concentrated close to the grounding lines (Dutrieux et al. 2014), but that it is nonnegligible elsewhere and the influence of that melt on the circulations we have explored remains to be investigated. Also, the importance of the missing surface forcing is important in nudging the stratification in otherwise more weakly stratified cases but has not been considered in this study.
Future open questions involve understanding how these dynamics interact with the evolving ice shelf and its impact on the cavity geometry and glacial melt rates. Additional considerations, including the variability of tidal and external currents, winds, accurate bathymetry, ice shelf geometry, and continuous stratification are all important next steps for the community to consider.
Acknowledgments
The authors thank Pierre St-Laurent for allowing the use of the open source code BEOM and insightful comments on the manuscript, Aviv Solodoch for comprehensive suggestions on the manuscript, Janine Schaffer for useful discussions and hydrographic data, Tore Hatterman for extensive feedback, and two anonymous reviewers for helpful comments. The model source code is available at www.nordet.net/beom.html. This material is based in part upon work supported by the National Science Foundation under Grant PLR-1543388.
APPENDIX A
Model Sensitivities
Although we have reduced the scope of our problem to heavily idealized cases, we can determine the sensitivity of the model results with three additional considerations: horizontal grid spacing, top topography, and quadratic friction. We find that in all cases the dynamics remain qualitatively unchanged, and our results may be expected to translate well to more realistic cavity configurations.
We find empirically that the transport, interface, velocity, and sill crossing location are not sensitive to horizontal grid spacings higher than

Resolution sensitivity plot of bottom layer potential vorticity [Eq. (11b)] for the case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Resolution sensitivity plot of bottom layer potential vorticity [Eq. (11b)] for the case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Resolution sensitivity plot of bottom layer potential vorticity [Eq. (11b)] for the case
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
In the main text, we restricted our attention to flat-topped rigid lids for simplicity. Therefore, these results are also applicable to the free surface exchange flows since the surface wave effects are negligible for our chosen parameters. However, we can test varying cases of top topographies, as shown in Fig. A2, which does not significantly impact the observable dynamics of the flows, except in extreme cases, which are possible since ice shelf drafts can vary by over hundreds of meters in domains with bathymetry comparable to ours. Top topography is defined using

Sensitivity plot of top and bottom layer thickness-weighted average potential vorticity (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1

Sensitivity plot of top and bottom layer thickness-weighted average potential vorticity (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
Sensitivity plot of top and bottom layer thickness-weighted average potential vorticity (
Citation: Journal of Physical Oceanography 49, 1; 10.1175/JPO-D-18-0076.1
We also tested longer domains and bathymetry for varying sill width scale
Finally, we tested the effect of a realistic quadratic friction with a dimensionless coefficient of 2 × 10−3, in contrast with the range of linear friction coefficients we have explored in the main text. With this quadratic friction, we observe the same dynamical regimes discussed in this study for equivalent values of friction depending on the specified pressure head. For example, with a pressure head of
APPENDIX B
Forward–Backward Time-Stepping in a Rigid Lid
We present a method for finding a convergent, accurate pressure field at the top surface for the forward–backward time-stepping scheme. A similar approach can also be taken for the generalized forward–backward time-stepping scheme.






































The velocities with the fully updated pressures must have zero depth-integrated divergence, satisfying the continuity equation at every grid point in the horizontal plane. Thus, the pressure gradient
APPENDIX C
Thickness-Weighted PV Balance
In our problem, the PV balance allows us to identify the dominant mechanisms that facilitate cross-sill exchanges in various parameter regimes.









APPENDIX D
Rotating Two-Layer Uniform PV Solution


















As prescribed pressure head increases, the analytical solution reaches a critical and eventually a supercritical state, which generally leads to PV adjustment processes that cause the breakdown of semigeostrophic theory and uniform PV becoming a poor assumption (Pratt and Armi 1990; Dalziel 1991). Therefore, there must be nonuniform PV effects, as otherwise the cross-channel velocity gradient would be highly supercritical for channel widths greater than
APPENDIX E
Rotating Uniform PV Layer-Grounding Theory




























REFERENCES
Aiki, H., X. Zhai, and R. Greatbatch, 2016: Energetics of the global ocean: The role of mesoscale eddies. Indo-Pacific Climate Variability and Predictability, S. K. Behera and T. Yamagata, Eds., World Scientific, 109–134.
Borenas, K., and J. Whitehead, 1998: Upstream separation in a rotating channel flow. J. Geophys. Res., 103, 7567–7578, https://doi.org/10.1029/97JC02548.
Cai, C., E. Rignot, D. Menemenlis, and Y. Nakayama, 2017: Observations and modeling of ocean-induced melt beneath Petermann Glacier Ice Shelf in northwestern Greenland. Geophys. Res. Lett., 44, 8396–8403, https://doi.org/10.1002/2017GL073711.
Dalziel, S., 1991: Two-layer hydraulics: A functional approach. J. Fluid Mech., 223, 135–163, https://doi.org/10.1017/S0022112091001374.
De Rydt, J., and G. Gudmundsson, 2016: Coupled ice shelf–ocean modeling and complex grounding line retreat from a seabed ridge. J. Geophys. Res. Earth Surf., 121, 865–880, https://doi.org/10.1002/2015JF003791.
De Rydt, J., P. R. Holland, P. Dutrieux, and A. Jenkins, 2014: Geometric and oceanographic controls on melting beneath Pine Island Glacier. J. Geophys. Res. Oceans, 119, 2420–2438, https://doi.org/10.1002/2013JC009513.
Dutrieux, P., and Coauthors, 2014: Strong sensitivity of Pine Island ice-shelf melting to climatic variability. Science, 343, 174–178, https://doi.org/10.1126/science.1244341.
Fedorov, A. V., and W. K. Melville, 1996: Hydraulic jumps at boundaries in rotating fluids. J. Fluid Mech., 324, 55–82, https://doi.org/10.1017/S0022112096007835.
Gill, A., 1977: The hydraulics of rotating-channel flow. J. Fluid Mech., 80, 641–671, https://doi.org/10.1017/S0022112077002407.
Griffies, S., and R. Hallberg, 2000: Biharmonic friction with a Smagorinsky-like viscosity for use in large-scale eddy-permitting ocean models. Mon. Wea. Rev., 128, 2935–2946, https://doi.org/10.1175/1520-0493(2000)128<2935:BFWASL>2.0.CO;2.
Gudmundsson, G. H., 2013: Ice-shelf buttressing and the stability of marine ice sheets. Cryosphere, 7, 647–655, https://doi.org/10.5194/tc-7-647-2013.
Gudmundsson, G. H., J. Krug, G. Durand, L. Favier, and O. Gagliardini, 2012: The stability of grounding lines on retrograde slopes. Cryosphere, 6, 1497–1505, https://doi.org/10.5194/tc-6-1497-2012.
Hallberg, R., and P. Rhines, 1996: Buoyancy-driven circulation in an ocean basin with isopycnals intersecting the sloping boundary. J. Phys. Oceanogr., 26, 913–940, https://doi.org/10.1175/1520-0485(1996)026<0913:BDCIAO>2.0.CO;2.
Helfrich, K. R., A. C. Kuo, and L. J. Pratt, 1999: Nonlinear Rossby adjustment in a channel. J. Fluid Mech., 390, 187–222, https://doi.org/10.1017/S0022112099005042.
Hogg, A. M., W. K. Dewar, P. Berloff, and M. L. Ward, 2011: Kelvin wave hydraulic control induced by interactions between vortices and topography. J. Fluid Mech., 687, 194–208, https://doi.org/10.1017/jfm.2011.344.
Holland, P. R., 2017: The transient response of ice shelf melting to ocean change. J. Phys. Oceanogr., 47, 2101–2114, https://doi.org/10.1175/JPO-D-17-0071.1.
Holland, P. R., A. Jenkins, and D. M. Holland, 2008: The response of ice shelf basal melting to variations in ocean temperature. J. Climate, 21, 2558–2572, https://doi.org/10.1175/2007JCLI1909.1.
Jacobs, S. S., A. Jenkins, C. F. Giulivi, and P. Dutrieux, 2011: Stronger ocean circulation and increased melting under Pine Island Glacier ice shelf. Nat. Geosci., 4, 519–523, https://doi.org/10.1038/ngeo1188.
Jenkins, A., P. Dutrieux, S. Jacobs, S. McPhail, J. Perrett, A. Webb, and D. White, 2010: Observations beneath Pine Island Glacier in West Antarctica and implications for its retreat. Nat. Geosci., 3, 468–472, https://doi.org/10.1038/ngeo890.
Joughin, I., R. B. Alley, and D. M. Holland, 2012: Ice-sheet response to oceanic forcing. Science, 338, 1172–1176, https://doi.org/10.1126/science.1226481.
Joughin, I., B. E. Smith, and B. Medley, 2014: Marine ice sheet collapse potentially under way for the Thwaites Glacier Basin, West Antarctica. Science, 344, 735–738, https://doi.org/10.1126/science.1249055.
Kimura, S., A. Jenkins, P. Dutrieux, A. Forryan, A. Naveira Garabato, and Y. Firing, 2016: Ocean mixing beneath Pine Island Glacier ice shelf, West Antarctica. J. Geophys. Res. Oceans, 121, 8496–8510, https://doi.org/10.1002/2016JC012149.
Konrad, H., L. Gilbert, S. L. Cornford, A. J. Payne, A. Hogg, A. Muir, and A. Shepherd, 2017: Uneven onset and pace of ice-dynamical imbalance in the Amundsen Sea Embayment, West Antarctica. Geophys. Res. Lett., 44, 910–918, https://doi.org/10.1002/2016GL070733.
Little, C. M., A. Gnanadesikan, and R. Hallberg, 2008: Large-scale oceanographic constraints on the distribution of melting and freezing under ice shelves. J. Phys. Oceanogr., 38, 2242–2255, https://doi.org/10.1175/2008JPO3928.1.
Lorenz, E. N., 1955: Available potential energy and the maintenance of the general circulation. Tellus, 7, 157–167, https://doi.org/10.1111/j.2153-3490.1955.tb01148.x.
Mayer, C., N. Reeh, F. Jung-Rothenhäusler, P. Huybrechts, and H. Oerter, 2000: The subglacial cavity and implied dynamics under Nioghalvfjerdsfjorden Glacier, NE-Greenland. Geophys. Res. Lett., 27, 2289–2292, https://doi.org/10.1029/2000GL011514.
Mouginot, J., E. Rignot, and B. Scheuchl, 2014: Sustained increase in ice discharge from the Amundsen Sea Embayment, West Antarctica, from 1973 to 2013. Geophys. Res. Lett., 41, 1576–1584, https://doi.org/10.1002/2013GL059069.
Nakayama, Y., R. Timmermann, M. Schröder, and H. Hellmer, 2014: On the difficulty of modeling circumpolar deep water intrusions onto the Amundsen Sea continental shelf. Ocean Modell., 84, 26–34, https://doi.org/10.1016/j.ocemod.2014.09.007.
Pratt, L. J., 1987: Rotating shocks in a separated laboratory channel flow. J. Phys. Oceanogr., 17, 483–491, https://doi.org/10.1175/1520-0485(1987)017<0483:RSIASL>2.0.CO;2.
Pratt, L. J., and L. Armi, 1990: Two-layer rotating hydraulics: Strangulation, remote and virtual controls. Pure Appl. Geophys., 133, 587–617, https://doi.org/10.1007/BF00876224.
Pratt, L. J., and J. A. Whitehead, 2007: Rotating Hydraulics: Nonlinear Topographic Effects in the Ocean and Atmosphere. Springer, 592 pp.
Pratt, L. J., K. R. Helfrich, and E. P. Chassignet, 2000: Hydraulic adjustment to an obstacle in a rotating channel. J. Fluid Mech., 404, 117–149, https://doi.org/10.1017/S0022112099007065.
Rignot, E., J. Mouginot, M. Morlighem, H. Seroussi, and B. Scheuchl, 2014: Widespread, rapid grounding line retreat of Pine Island, Thwaites, Smith, and Kohler glaciers, West Antarctica, from 1992 to 2011. Geophys. Res. Lett., 41, 3502–3509, https://doi.org/10.1002/2014GL060140.
Sadourny, R., 1975: The dynamics of finite-difference models of the shallow-water equations. J. Atmos. Sci., 32, 680–689, https://doi.org/10.1175/1520-0469(1975)032<0680:TDOFDM>2.0.CO;2.
Salmon, R., 2002: Numerical solution of the two-layer shallow water equations with bottom topography. J. Mar. Res., 60, 605–638, https://doi.org/10.1357/002224002762324194.
Schaffer, J., 2017: Ocean impact on the 79 North Glacier, Northeast Greenland. Ph.D. thesis, University of Bremen, 206 pp.
Schodlok, M., D. Menemenlis, E. Rignot, and M. Studinger, 2012: Sensitivity of the ice-shelf/ocean system to the sub-ice-shelf cavity shape measured by NASA IceBridge in Pine Island Glacier, West Antarctica. Ann. Glaciol., 53, 156–162, https://doi.org/10.3189/2012AoG60A073.
Schoof, C., 2007: Ice sheet grounding line dynamics: Steady states, stability, and hysteresis. J. Geophys. Res., 112, F03S282, https://doi.org/10.1029/2006JF000664.
Seroussi, H., Y. Nakayama, E. Larour, D. Menemenlis, M. Morlighem, E. Rignot, and A. Khazendar, 2017: Continued retreat of Thwaites Glacier, West Antarctica, controlled by bed topography and ocean circulation. Geophys. Res. Lett., 44, 6191–6199, https://doi.org/10.1002/2017GL072910.
Shchepetkin, A. F., and J. C. McWilliams, 1998: Quasi-monotone advection schemes based on explicit locally adaptive dissipation. Mon. Wea. Rev., 126, 1541–1580, https://doi.org/10.1175/1520-0493(1998)126<1541:QMASBO>2.0.CO;2.
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.
Stern, M., 1974: Comment on rotating hydraulics. Geophys. Fluid Dyn., 6, 127–130, https://doi.org/10.1080/03091927409365791.
St-Laurent, P., 2018: Back of Envelope Ocean Model (BEOM). Accessed 1 April 2018, www.nordet.net/beom.html.
St-Laurent, P., J. M. Klinck, and M. S. Dinniman, 2015: Impact of local winter cooling on the melt of Pine Island Glacier, Antarctica. J. Geophys. Res. Oceans, 120, 6718–6732, https://doi.org/10.1002/2015JC010709.
Stommel, H., 1948: The westward intensification of wind-driven ocean currents. Trans. Amer. Geophys., 29, 202–206, https://doi.org/10.1029/TR029i002p00202.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 745 pp.
Whitehead, J. A., A. Leetmaa, and R. Knox, 1974: Rotating hydraulics of strait and sill flows. Geophys. Fluid Dyn., 6, 101–125, https://doi.org/10.1080/03091927409365790.