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

212 38 CHAPTER 6. GYRE CIRCULATION 6.2 The Ekman Layer tel-00545911, version 1 - 13 Dec 2010 Strictly speaking, this section belongs to chapter 10 because it deals with how wind forcing penetrates to the deep ocean, but it just happens that we need to know Ekman theory to continue our investigation of “Forced Dynamics.” Ekman’s theory of the adjustment of a fluid in a rotating frame to an equilibrium when subject to wind forcing, is probably the most cited and most misunderstood theory of ocean dynamics. To elucidate this Ekman layer dynamics we will advance in small steps, emphasizing the physical understanding of the process, without neglecting the mathematical derivation. Suppose we have an infinitely deep layer of a homogeneous fluid subject to wind forcing τ x , constant in time and space, at its surface that is acting in the x-direction. The flow is independent of x,y as the forcing has no variations in these variables and as there are no boundaries. But the flow depends on the vertical coordinate z. In this case the vertical velocity w vanishes everywhere due to the divergence free condition, eq. (5.4). The Navier-Stokes equations (5.1) – (5.4), in a rotating frame, then simplify to: with the boundary conditions: ∂ t u(z,t) − fv(z,t) = ν∂ zz u(z,t) (6.9) ∂ t v(z,t) + fu(z,t) = ν∂ zz v(z,t) (6.10) ν∂ z u(0) = τ x ; ∂ z v(0) = 0, (6.11) lim ∂ zu(z) = z→−∞ lim ∂ zv(z) = 0. (6.12) z→−∞ The surface boundary condition (6.11) represents the vertical gradient of the horizontal velocity due to wind stress, while we suppose no frictional forces at the (far away) bottom of the Ekman layer. 6.2.1 Ekman Transport (one layer) To further simplify the problem we consider the transport U(t) = ∫ 0 ∫ 0 −H −H u(z,t)dz and V (t) = u(z,t)dz of the whole fluid column. Please note, that these variables depend only on time and we have neglected all vertical structure in the problem. This can be easily done in the present problem as eqs. (6.9), (6.10) and the boundary conditions (6.11), (6.12) are linear. Integrating the right hand side of eq. (6.9) we have ∫ 0 ν∂ −H zzu(z,t)dz = ν∂ z u(0) = τ x . When we further neglect friction at the bottom of the fluid layer eq. (6.9), (6.10) and boundary conditions (6.12) read: ∂ t U(t) − fV (t) = τ x (6.13) ∂ t V (t) + fU(t) = 0 (6.14) We now like to consider the spin up of an Ekman transport initially at rest. In the nonrotating case (f = 0) we have the solution: U(t) = τ x t (6.15) V (t) = 0, (6.16) so the fluid constantly accelerates in the x-direction and no stationary state is reached!

213 6.2. THE EKMAN LAYER 39 In the rotating case (f ≠ 0) the solution is given by: U(t) = τx f sin(ft) (6.17) V (t) = τx f (cos(ft) − 1) (6.18) Initially the solution behaves as in the non rotating case, that is, it accelerates in the x-direction with an acceleration given by τ x . But in the rotating case eqs. (6.13) and (6.14) also have the stationary (time-independent) solution: tel-00545911, version 1 - 13 Dec 2010 U = 0; V = − τ x f , (6.19) which has no counter part in the non-rotating case. This solution is a force balance between the Coriolis force and the wind stress. The depth averaged Ekman transport is at 90 o to the right of the wind force as this is the only possibility for the Coriolis force to balance the wind stress. ✻ wind stress Coriolis force ❄ Ekman ✲ transport Figure 6.3: Depth averaged Ekman transport The solutions for the rotating case given in eqs. 6.17 and 6.18 are in fact a sum of the stationary solution plus inertial oscillation. When friction is included the oscillations will be damped and the transport will converge towards a (modified) stationary solution. Exercise 38: What is the energetics of the Ekman transport? Exercise 39: Does the Ekman transport depend on the viscosity? It is the Ekman transport, and only the Ekman transport, that determines the influence of the wind forcing on the deep ocean. For completeness we will discuss in the next subsection the vertical structure of the Ekman dynamics. 6.2.2 The Ekman Spiral We start this section with two instructive exercises. Exercise 40: What happens when we include bottom (Rayleigh) friction in eqs. 6.13 and 6.14? The stationary solution is governed by: −fV = −rU + τ x (6.20) fU = −rV (6.21) and we obtain the solution: U = V = r r 2 + f 2τ x (6.22) −f r 2 + f 2τ x. (6.23)

212<br />

38 CHAPTER 6. GYRE CIRCULATION<br />

6.2 The Ekman Layer<br />

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

Strictly speaking, this section belongs to chapter 10 because it <strong>de</strong>als with how wind forcing<br />

pen<strong>et</strong>rates to the <strong>de</strong>ep ocean, but it just happens that we need to know Ekman theory to<br />

continue our investigation of “Forced Dynamics.”<br />

Ekman’s theory of the adjustment of a fluid in a rotating frame to an equilibrium when<br />

subject to wind forcing, is probably the most cited and most misun<strong>de</strong>rstood theory of ocean<br />

dynamics. To elucidate this Ekman layer dynamics we will advance in small steps, emphasizing<br />

the physical un<strong>de</strong>rstanding of the process, without neglecting the mathematical <strong>de</strong>rivation.<br />

Suppose we have an infinitely <strong>de</strong>ep layer of a homogeneous fluid subject to wind forcing<br />

τ x , constant in time and space, at its surface that is acting in the x-direction. The flow is<br />

in<strong>de</strong>pen<strong>de</strong>nt of x,y as the forcing has no variations in these variables and as there are no<br />

boundaries. But the flow <strong>de</strong>pends on the vertical coordinate z. In this case the vertical velocity<br />

w vanishes everywhere due to the divergence free condition, eq. (5.4). The Navier-Stokes<br />

equations (5.1) – (5.4), in a rotating frame, then simplify to:<br />

with the boundary conditions:<br />

∂ t u(z,t) − fv(z,t) = ν∂ zz u(z,t) (6.9)<br />

∂ t v(z,t) + fu(z,t) = ν∂ zz v(z,t) (6.10)<br />

ν∂ z u(0) = τ x ; ∂ z v(0) = 0, (6.11)<br />

lim ∂ zu(z) =<br />

z→−∞<br />

lim ∂ zv(z) = 0. (6.12)<br />

z→−∞<br />

The surface boundary condition (6.11) represents the vertical gradient of the horizontal velocity<br />

due to wind stress, while we suppose no frictional forces at the (far away) bottom of the Ekman<br />

layer.<br />

6.2.1 Ekman Transport (one layer)<br />

To further simplify the problem we consi<strong>de</strong>r the transport U(t) = ∫ 0<br />

∫ 0<br />

−H<br />

−H<br />

u(z,t)dz and V (t) =<br />

u(z,t)dz of the whole fluid column. Please note, that these variables <strong>de</strong>pend only on time<br />

and we have neglected all vertical structure in the problem. This can be easily done in the<br />

present problem as eqs. (6.9), (6.10) and the boundary conditions (6.11), (6.12) are linear.<br />

Integrating the right hand si<strong>de</strong> of eq. (6.9) we have ∫ 0<br />

ν∂ −H zzu(z,t)dz = ν∂ z u(0) = τ x . When<br />

we further neglect friction at the bottom of the fluid layer eq. (6.9), (6.10) and boundary<br />

conditions (6.12) read:<br />

∂ t U(t) − fV (t) = τ x (6.13)<br />

∂ t V (t) + fU(t) = 0 (6.14)<br />

We now like to consi<strong>de</strong>r the spin up of an Ekman transport initially at rest. In the nonrotating<br />

case (f = 0) we have the solution:<br />

U(t) = τ x t (6.15)<br />

V (t) = 0, (6.16)<br />

so the fluid constantly accelerates in the x-direction and no stationary state is reached!

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

Saved successfully!

Ooh no, something went wrong!