Etudes et évaluation de processus océaniques par des hiérarchies ...

Etudes et évaluation de processus océaniques par des hiérarchies ... Etudes et évaluation de processus océaniques par des hiérarchies ...

tel.archives.ouvertes.fr
from tel.archives.ouvertes.fr More from this publisher
27.12.2013 Views

132 CHAPITRE 4. ETUDES DE PROCESSUS OCÉANOGRAPHIQUES 552 Ocean Dynamics (2009) 59:551–563 tel-00545911, version 1 - 13 Dec 2010 momentum are neglected, rotating gravity currents flow along the inclined ocean floor changing neither depth nor composition. Turbulent fluxes of tracers, that is mixing, entrainment and/or detrainment, determine the change of the water mass. Together with the turbulent fluxes of momentum at the floor and interface of the gravity current, they govern the pathway and the composition of the gravity current. Oceanic gravity currents are thus subject to three forces: (1) the buoyancy force pointing downslope, (2) the Coriolis force directed at a right angle to the direction of motion and (3) frictional forces due to dissipative fluxes of mass and momentum (see Fig. 7 in the Appendix). The first two forces are linear, the total buoyancy force depends linearly on the total buoyancy anomaly and the total Coriolis force is a linear function of the transport of the gravity current. The above shows that the answer and problem of the evolution of a gravity current lies in the determination of the laws and parameters of its dissipative fluxes of momentum and mass. For an introduction to the dynamics of oceanic gravity currents, I refer the reader to Griffiths (1986). In the past 30 years, the study of streamtube models has greatly increased our understanding of the dynamics of oceanic gravity currents. In streamtube models, only average (bulk) values of the dynamical variables are considered at every cross section along the path of the gravity current (see Smith 1975, 1977; Killworth 1977, 2001; Price and Baringer 1994; Baringer and Price 1997; Emms 1998). Due to their linear nature, the integrated buoyancy and Coriolis force can be calculated. The nonlinear frictional forces and the turbulent fluxes, perturbing the geostrophic equilibrium, have to be parametrised. The evolution of the gravity current can then be studied as a function of the parametrisations and parameter values employed. Integrations of these semi-analytical streamtube models along a path of varying topographic slope are performed and the parameter values are adjusted in such a way that the pathway and water mass characteristics compare favourably to observations. This adjustment procedure, however, does not always lead to a deeper understanding of the physical processes involved (see Emms 1998). The implementation of streamtube models or their physics in large-scale ocean general circulation models is difficult. Although progress has been made in better representing gravity currents in large-scale ocean general circulation models through bottom boundary layers (see, e.g. Killworth and Edwards 1999; Price and Young 1998; Wu et al. 2007), their dynamics is still a weak point of today’s numerical models of ocean circulation. Many efforts have been undertaken to determine the turbulent fluxes from observations, laboratory experiments and numerical simulations, but no definite conclusion has been reached, no generally accepted parametrisation is available and no generally accepted parameter values have been obtained so far. As an example, a key parameter is the entrainment rate; there is, however, no definite answer even about its sign in rotating gravity currents (Ezer 2005; Legg et al. 2006). Killworth (2001) states that “entrainment should only occur over limited regions, with detrainment elsewhere”, and MacCready notes that “[...] the majority of entrainment with overlying waters occurs close to the overflow sill [...], and is largely negligible there after”. Furthermore, Ermanyuk and Gavrilov (2007) found in laboratory experiments on non-rotating gravity currents that the dissipative processes at the interface have a negligible role compared to those due to bottom friction. The purpose of the present work is to determine the basic processes and structure of oceanic gravity currents in an idealised configuration as a testing ground for present and future models and parametrisations in basin scale models. To this end, I will numerically integrate the (nonhydrostatic) Navier–Stokes equations. With respect to streamtube models, I will not try to adjust parameter values so that they compare favourably to observation, but I will instead test the hypotheses they are based on by comparing their assumptions with our results from the Navier–Stokes model of highly idealised gravity currents. I show that gravity currents actually consist of two parts and that this two-part structure is key to understanding and modelling their dynamics. In his pioneering paper on the dynamics of viscousrotating-gravity currents, Smith (1977) stated, concerning their structure, “[...] a complete solution remains inaccessible” (Section 3, lines 7–8). I will here demonstrate that the structure, evolving in time, of idealised gravity currents can be obtained by solving a onedimensional heat equation. Results on the basic structure of idealised oceanic gravity currents are also the starting point for the large-scale, three-dimensional instability analysis leading to the formation of coherent cyclonic structures (Meacham and Stephens 2001). The important dynamics of large-scale instability is not considered in this work. The model is introduced in the next section, results are presented in Section 3 and they are discussed in Section 4, where I also consider the consequences of our results for the representation of oceanic gravity currents in ocean general circulation models (OGCMs).

4.7. ON THE BASIC STRUCTURE OF OCEANIC GRAVITY CURRENTS 133 Ocean Dynamics (2009) 59:551–563 553 tel-00545911, version 1 - 13 Dec 2010 2 Idealised oceanic gravity current 2.1 The physical problem considered In the experiments presented here, I use an idealised geometry, considering an infinitely long gravity current on an inclined plane with constant slope, and I do not allow for variations in the long-stream direction. A similar geometry was investigated by Ezer and Weatherly (1990). I thus consider only the dynamics of a vertical slice perpendicular to the geostrophic flow direction of the gravity current. Please note that such simplified geometry inhibits large-scale instability and the formation of cyclones and other large-scale features. This is beneficial to our goal of studying the small-scale fluxes in gravity currents. The simplified geometry also filters out small-scale processes and instabilities in the y direction. Instead of a gravity current descending in space, along the direction of propagation, it descends in time (this strategy was also used by MacCready 1994 and others). Such descent is also investigated when the gravity current is initially homogeneous in the longslope direction, as in laboratory experiments where the dense fluid is injected axisymmetrically on a cone structure (Shapiro and Zatsepin 1997; Sutherland et al. 2004); semi-analytical calculations of the instability of a rotating gravity current (Meacham and Stephens 2001). The results can be compared to the observations of the transverse structure of a oceanic gravity current by Umlauf et al. (2007). The local dynamics will not differ from gravity currents descending in space as long as the large-scale descent is slow in space and time compared to the local turbulent dynamics, which is definitely the case when the dynamics is close to a geostrophic equilibrium. Such type of model is referred to as 2.5- dimensional, as it is three-dimensional, but the variables have no dependence on the stream-wise direction (the y direction in this case). I like to emphasise that none of the three components of the velocity vector are trivial and that the main (geostrophic) transport is in the y direction. In observations and laboratory and numerical experiments, large-scale instabilities are observed for a wide range of parameter values. All the studies, that I am aware of, include a geostrophic adjustment process, which also destabilises the current. In laboratory experiments where the gravity current was injected close to a geostrophically adjusted state, these large-scale instabilities developed only very slowly (Wirth and Sommeria, unpublished manuscript), long time after the 2.5-dimensional dynamics, studied here, had started evolving. A geostrophically adjusted state also forms the starting point of investigations concerning the stability of oceanic gravity currents and the dynamics of streamtube models. The gravity currents considered here are on an inclined plane of a constant slope of one degree, the initial profile of its vertical extension above the inclined ocean floor z = h(x) = max{H − x 2 /λ, 0} has a parabolic shape with H = 200 m and λ = 5. · 10 5 m leading to a gravity current that is 200 m high and L = 20 km large. The values used in the experiment are typical for oceanic gravity currents (see, e.g. Smith 1977; Price and Baringer 1994; Killworth 2001). If the gravity current is initially geostrophically adjusted, the velocity components are given by: u G = 0; v G = g′ (∂ x h + tan α) ; w G = 0. (1) f This geostrophic velocity, also called the Nof speed (Nof 1983), is due to the balance between the Coriolis force and the buoyancy force in the downslope direction. Outside the gravity current, the water is initially at rest. Multiplying Eq. 1 by the thickness and integrating across the gravity current shows that the average geostrophic speed of the gravity current is given by v G = (g ′ /f) tan α. The Coriolis parameter is f = 1.0313 · 10 −4 s −1 corresponding to the earth rotation at mid-latitude. The reduced gravity is g ′ = g(ρ gc − ρ 0 )/ρ 0 = T · 2. · 10 −4 K −1 9.8065 m s −2 , where T is the temperature difference between the ambient fluid and the gravity current, a constant thermal expansion coefficient of 2 · 10 −4 K −1 and a gravitational acceleration of 9.8065 m s −2 are used. The temperature difference between the gravity current and the surrounding water ranges from 0.25 to 1.5 K. The values of the average geostrophic speed are given in Table 1. The vertical friction coefficient is ν v = 10 −3 m 2 s −1 , leading to an Ekman layer thickness δ = √ 2ν v /f ≈ 4.4 m. The dynamics depends on the six independent parameters (α, g ′ , f, H, L, ν v ); four independent nondimensional numbers can be obtained: (1) the slope of the inclined plane tan α, (2) the Richardson number Ri = g ′ H/v G 2 = H f 2 /(g ′ tan 2 α), (3) the Ekman number Ek = (δ/H) 2 = 2ν v /( f H 2 ) = 4.84 · 10 −4 comparing the frictional to the Coriolis force and (4) the ratio L/L D , the width of the gravity current divided by the Rossby radius L D = √ g ′ H/f. In the experiments presented here, I choose to systematically vary the Richardson number, identified as a key parameter (see, e.g. Price and Baringer 1994; Killworth 2001), by varying the density (temperature) difference. The Froude number, F = Ri −1/2 , gives the geostrophic speed divided by the speed of the shallow water gravity wave.

4.7. ON THE BASIC STRUCTURE OF OCEANIC GRAVITY CURRENTS 133<br />

Ocean Dynamics (2009) 59:551–563 553<br />

tel-00545911, version 1 - 13 Dec 2010<br />

2 I<strong>de</strong>alised oceanic gravity current<br />

2.1 The physical problem consi<strong>de</strong>red<br />

In the experiments presented here, I use an i<strong>de</strong>alised<br />

geom<strong>et</strong>ry, consi<strong>de</strong>ring an infinitely long gravity current<br />

on an inclined plane with constant slope, and I do not<br />

allow for variations in the long-stream direction. A similar<br />

geom<strong>et</strong>ry was investigated by Ezer and Weatherly<br />

(1990). I thus consi<strong>de</strong>r only the dynamics of a vertical<br />

slice perpendicular to the geostrophic flow direction<br />

of the gravity current. Please note that such simplified<br />

geom<strong>et</strong>ry inhibits large-scale instability and the<br />

formation of cyclones and other large-scale features.<br />

This is beneficial to our goal of studying the small-scale<br />

fluxes in gravity currents. The simplified geom<strong>et</strong>ry also<br />

filters out small-scale processes and instabilities in the<br />

y direction. Instead of a gravity current <strong>de</strong>scending in<br />

space, along the direction of propagation, it <strong>de</strong>scends<br />

in time (this strategy was also used by MacCready 1994<br />

and others). Such <strong>de</strong>scent is also investigated when the<br />

gravity current is initially homogeneous in the longslope<br />

direction, as in laboratory experiments where<br />

the <strong>de</strong>nse fluid is injected axisymm<strong>et</strong>rically on a cone<br />

structure (Shapiro and Zatsepin 1997; Sutherland <strong>et</strong> al.<br />

2004); semi-analytical calculations of the instability of a<br />

rotating gravity current (Meacham and Stephens 2001).<br />

The results can be com<strong>par</strong>ed to the observations of<br />

the transverse structure of a oceanic gravity current by<br />

Umlauf <strong>et</strong> al. (2007). The local dynamics will not differ<br />

from gravity currents <strong>de</strong>scending in space as long as the<br />

large-scale <strong>de</strong>scent is slow in space and time com<strong>par</strong>ed<br />

to the local turbulent dynamics, which is <strong>de</strong>finitely<br />

the case when the dynamics is close to a geostrophic<br />

equilibrium. Such type of mo<strong>de</strong>l is referred to as 2.5-<br />

dimensional, as it is three-dimensional, but the variables<br />

have no <strong>de</strong>pen<strong>de</strong>nce on the stream-wise direction<br />

(the y direction in this case). I like to emphasise that<br />

none of the three components of the velocity vector are<br />

trivial and that the main (geostrophic) transport is in<br />

the y direction.<br />

In observations and laboratory and numerical experiments,<br />

large-scale instabilities are observed for a<br />

wi<strong>de</strong> range of <strong>par</strong>am<strong>et</strong>er values. All the studies, that I<br />

am aware of, inclu<strong>de</strong> a geostrophic adjustment process,<br />

which also <strong>de</strong>stabilises the current. In laboratory experiments<br />

where the gravity current was injected close<br />

to a geostrophically adjusted state, these large-scale<br />

instabilities <strong>de</strong>veloped only very slowly (Wirth and<br />

Sommeria, unpublished manuscript), long time after<br />

the 2.5-dimensional dynamics, studied here, had started<br />

evolving. A geostrophically adjusted state also forms<br />

the starting point of investigations concerning the stability<br />

of oceanic gravity currents and the dynamics of<br />

streamtube mo<strong>de</strong>ls.<br />

The gravity currents consi<strong>de</strong>red here are on an inclined<br />

plane of a constant slope of one <strong>de</strong>gree, the<br />

initial profile of its vertical extension above the inclined<br />

ocean floor z = h(x) = max{H − x 2 /λ, 0} has a<br />

<strong>par</strong>abolic shape with H = 200 m and λ = 5. · 10 5 m<br />

leading to a gravity current that is 200 m high and<br />

L = 20 km large. The values used in the experiment<br />

are typical for oceanic gravity currents (see, e.g. Smith<br />

1977; Price and Baringer 1994; Killworth 2001). If the<br />

gravity current is initially geostrophically adjusted, the<br />

velocity components are given by:<br />

u G = 0; v G = g′ (∂ x h + tan α)<br />

; w G = 0. (1)<br />

f<br />

This geostrophic velocity, also called the Nof speed<br />

(Nof 1983), is due to the balance b<strong>et</strong>ween the Coriolis<br />

force and the buoyancy force in the downslope<br />

direction. Outsi<strong>de</strong> the gravity current, the water is initially<br />

at rest. Multiplying Eq. 1 by the thickness and<br />

integrating across the gravity current shows that the<br />

average geostrophic speed of the gravity current is<br />

given by v G = (g ′ /f) tan α. The Coriolis <strong>par</strong>am<strong>et</strong>er is<br />

f = 1.0313 · 10 −4 s −1 corresponding to the earth rotation<br />

at mid-latitu<strong>de</strong>. The reduced gravity is g ′ = g(ρ gc −<br />

ρ 0 )/ρ 0 = T · 2. · 10 −4 K −1 9.8065 m s −2 , where T is<br />

the temperature difference b<strong>et</strong>ween the ambient fluid<br />

and the gravity current, a constant thermal expansion<br />

coefficient of 2 · 10 −4 K −1 and a gravitational acceleration<br />

of 9.8065 m s −2 are used. The temperature difference<br />

b<strong>et</strong>ween the gravity current and the surrounding<br />

water ranges from 0.25 to 1.5 K. The values of the<br />

average geostrophic speed are given in Table 1. The<br />

vertical friction coefficient is ν v = 10 −3 m 2 s −1 , leading<br />

to an Ekman layer thickness δ = √ 2ν v /f ≈ 4.4 m.<br />

The dynamics <strong>de</strong>pends on the six in<strong>de</strong>pen<strong>de</strong>nt <strong>par</strong>am<strong>et</strong>ers<br />

(α, g ′ , f, H, L, ν v ); four in<strong>de</strong>pen<strong>de</strong>nt nondimensional<br />

numbers can be obtained: (1) the slope<br />

of the inclined plane tan α, (2) the Richardson number<br />

Ri = g ′ H/v G 2 = H f 2 /(g ′ tan 2 α), (3) the Ekman number<br />

Ek = (δ/H) 2 = 2ν v /( f H 2 ) = 4.84 · 10 −4 com<strong>par</strong>ing<br />

the frictional to the Coriolis force and (4) the ratio<br />

L/L D , the width of the gravity current divi<strong>de</strong>d by<br />

the Rossby radius L D = √ g ′ H/f. In the experiments<br />

presented here, I choose to systematically vary the<br />

Richardson number, i<strong>de</strong>ntified as a key <strong>par</strong>am<strong>et</strong>er (see,<br />

e.g. Price and Baringer 1994; Killworth 2001), by varying<br />

the <strong>de</strong>nsity (temperature) difference. The Frou<strong>de</strong><br />

number, F = Ri −1/2 , gives the geostrophic speed divi<strong>de</strong>d<br />

by the speed of the shallow water gravity wave.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!