<<

A STEADY STRATIFIED PURELY AZIMUTHAL FLOW REPRESENTING THE ANTARCTIC CIRCUMPOLAR CURRENT

CALIN IULIAN MARTIN

Faculty of Mathematics, University of Vienna, Austria E-mail address: [email protected]

RONALD QUIRCHMAYR∗

Department of Mathematics, KTH Royal Institute of Technology, Sweden E-mail address: [email protected]

Abstract. We construct an explicit steady stratified purely azimuthal flow for the governing equations of geophysical fluid dynamics. These equations are con- sidered in a setting that applies to the Antarctic Circumpolar Current, accounting for and forcing terms.

Keywords: Antarctic Circumpolar Current, variable density, azimuthal flows, eddy viscosity, geophysical fluid dynamics. Mathematics Subject Classification: Primary: 35Q31, 35Q35. Secondary: 35Q86.

1. Introduction Thorough analytical investigations of exact solutions to the fully nonlinear governing equations of geophysical fluid dynamics (GFD) represent an extensive and active research area, which was initiated by Constantin [1–4] and Constantin and Johnson [5–8]. Following this approach, we construct a unidirectional flow satisfying GFD considered in the so-called f-plane at the 45th parallel , enhanced with an eddy viscosity term and a forcing term, and equipped with appropriate boundary conditions. We propose this specific flow for representing the gross dynamics of the arXiv:2010.09389v1 [physics.flu-dyn] 19 Oct 2020 Antarctic Circumpolar Current (ACC)—the World’s longest and strongest current. ACC has no continental barriers: it encircles along a 23 000 km path around the polar axis towards East at latitudes between 40° to 60°, see Fig. 1. It thereby links the Atlantic, Pacific and Indian making it the most impor- tant oceanic current in the ’s climate system. ACC has a rich and complicated

∗Corresponding author. 1 2 C. I. MARTIN AND R. QUIRCHMAYR structure. Many factors contribute to its complex behavior—the most important dri- ver are the strong westerly in the region. In addition to that there exist mesoscale eddies of a size up to 100 km, which transport the -induced surface to the bottom and also enable meridional transport; there are sharp changes in water density due to variations in temperature and salinity—known as fronts or jets—located at ACC’s boundaries (see Fig. 1); ACC is strongly con- strained by the bottom topography; there are observed variations in time such as the Antarctic Circumpolar Wave; etc. We refer to [9, 11, 12, 16, 18, 19, 24, 25, 27, 28] for further information about the geophysical aspects and modeling as well as ob- servational data and simulations for ACC. From an analytical perspective one is forced to largely, yet reasonably, simplify the geophysical scenario to obtain a tractable model, which—in the ideal case—exhibits exact and explicit solutions opening the path for an in-depth analysis. Thus we do not account for all of the before mentioned phenomena, but assume a steady flow in purely azimuthal direction, which is vertically bounded by a flat bottom and a flat ocean surface. By considering Euler’s equation of motion in the f-plane, we obtain a valid approximation of the Coriolis effects close to the 45th parallel south; in this way the Earth’s curvature is neglected and no boundaries in the meridional direction are assumed. To account for the transportation effects of mesoscale eddies we equip the system with an eddy viscosity term; furthermore we include a forcing term to ensure the dynamical balance of the flow. Both pressure and wind stress are prescribed on the ocean surface; a no-slip boundary condition is assumed for the ocean bed. A similar setting has recently been considered in [22], where an explicit solution in terms of a given viscosity function was presented. In this note at hand we extend these results to stratified flows, i.e. we do account for variations of the water density (with depth and latitude). The established explicit solution is an analytic function of both the viscosity function and the density distribution. A collection of numerous recent related analytical studies can be found in [7,13,14,17,20,22] and the references therein.

2. Model under study We introduce the governing equations for geophysical ocean flows that set the basis for our study: we take into account the effects of the Earth’s rotation on the fluid body by choosing a rotating framework with the origin at a point on the Earth’s surface. Accordingly, we will work with Cartesian coordinates (x, y, z), where x de- notes the direction of increasing longitude, y is the direction of increasing latitude and z represents the local vertical, respectively. Denoting with t the time vari- able, and with (u(x, y, z, t), v(x, y, z, t), w(x, y, z, t)) the velocity field, the governing equations for inviscid and incompressible geophysical ocean flows at latitude φ are A STEADY STRATIFIED PURELY AZIMUTHAL FLOW REPRESENTING THE ACC 3

Figure 1. Variations in density and the mean eastward current speed. The black lines in the image represent the fronts of ACC. Im- age credit: Hellen Phillips (Senior Research Fellow, Institute for Ma- rine and Antarctic Studies, University of ), Benoit Legresy (CSIRO) and Nathan Bindoff (Professor of , Institute for Marine and Antarctic Studies, University of Tasmania)

(cf. [3,21,26]) the Euler equations 1 u + uu + vu + wu + 2Ω(w cos φ − v sin φ) = − P , t x y z ρ x 1 v + uv + vv + wv + 2Ωu sin φ = − P , (2.1) t x y z ρ y 1 w + uw + vw + ww − 2Ωu cos φ = − P − g, t x y z ρ z and the equation of mass conservation div(ρu) = 0, (2.2) where u = (u, v, w). Here, P = P (x, y, z, t) denotes the pressure field, ρ = ρ(y, z) serves a (prescribed) density distribution to account for density gradients, mainly caused by variations of salinity and temperature, Ω ≈ 7.29 · 10−5 rad s−1 is the 4 C. I. MARTIN AND R. QUIRCHMAYR

Earth’s (constant) rotational speed around the polar axis toward the East, and g ≈ 9.82 m/s2 denotes the gravitational constant. The governing equations (2.1)– (2.2) hold throughout the fluid domain, which lies between the rigid flat bed at z = −d (d is the constant water depth) and above the flat surface situated at z = 0. Our aim is to derive exact formulas for purely azimuthal flows (i.e. v = w = 0) in the region of the ACC (i.e. we take φ = −π/4, which corresponds to the 45th parallel south). Furthermore we incorporate the transfer of the wind-generated surface stress to the bottom, which is due to the presence of mesoscale eddies, by adding a viscosity term of the form (νuz)z to the right hand side of the first equation in (2.1); the coefficient ν = ν(z) is a smooth function of depth being strictly greater than some positive constant, see [16]. The classical model of uniform eddy viscosity is due to [23]. We follow the more realistic approach with a depth dependent viscosity function as it was introduced in [10]. Finally, we include a forcing term F = F (y, z) to guarantee non-trivial solutions, cf. [7,15,22]. According to the previous considerations, we will consider the following set of equations governing the ACC as a geophysical flow model in the purely azimuthal direction: 1 0 = − P + (νu ) (2.3) ρ x z z √ 1 − 2Ωu = − P + F (2.4) ρ y √ 1 − 2Ωu = − P − g (2.5) ρ z

ux = 0 (2.6) and are valid within the fluid domain D := R2 × [−d, 0] ⊂ R3. The equations of motion (2.3)-(2.6) are supplemented by the following boundary conditions: u = 0 on z = −d, (2.7)

P = Patm on z = 0, (2.8)

τ = τ0 on z = 0, (2.9) where τ(y, z) := ρνuz represents the wind stress; i.e. we assume a no-slip bottom and constant pressure as well as wind stress at the surface.

3. Explicit solution

Theorem 3.1. The solution (u, P ) of system (2.3)–(2.6) with boundary conditions (2.7)–(2.9) is given by τ Z z dζ u(y, z) = 0 (3.1) ρ(y, 0) −d ν(ζ) A STEADY STRATIFIED PURELY AZIMUTHAL FLOW REPRESENTING THE ACC 5 and √ Z z Z z P (y, z) = 2Ω ρ(y, ζ)u(y, ζ) dζ − g ρ(y, ζ) dζ + Patm, (3.2) 0 0 for (y, z) ∈ R×[−d, 0]. Furthermore, the forcing term F can be recovered from (2.4) by means of (3.1) and (3.2). Proof. Utilizing (2.6) we differentiate by x in (2.3)-(2.5) and obtain that

∇Px = 0 in D. (3.3)

From the condition P = Patm on the surface z = 0 we obtain that Px = 0 on z = 0. Thus, from (3.3) we see that Px = 0 in D.

Hence, equation (2.3) becomes (νuz)z = 0, which implies (making also use of (2.6)) that there is a funtion y 7→ B(y) such that

νuz = B(y) in D. The latter equation and condition (2.9) imply that τ ν(0)u (y, 0) = 0 = B(y) for all y. z ρ(y, 0)

τ0 Therefore, ν(z)uz(y, z) = ρ(y,0) for all y and for all z ∈ [−d, 0]. Thus, by means of the bottom boundary condition (2.7) we infer that u satisfies (3.1). Integrating with respect to z in (2.5) we obtain that √ Z z Z z P (y, z) = 2Ω ρ(y, ζ)u(y, ζ) dζ − g ρ(y, ζ) dζ + C(y), −d −d where the integration constant C(y) is determined by (2.8) and satisfies √ Z 0 Z 0 C(y) = Patm − 2Ω ρ(y, z)u(y, z) dz + g ρ(y, z) dz. −d −d Therefore, P satisfies (3.2) for all (y, z) ∈ R × [−d, 0].  Remark 3.2. One immediate consequence of (3.1) is that the vorticity vector as- sociated with the flow (3.1) is (0, uz, −uy) has a non-vanishing second and third component. This represents a marked difference, if compared with the case of ho- mogeneous flows (considered in [22]) where only the middle component uz survives, the first and the third being zero because of the lack of y dependence of ρ. Thus, allowing for significant variations in density leads to solutions that exhibit substan- tial shear not only in the vertical direction but also in the latitudinal direction, as well.

Acknowledgments C. I. Martin would like to acknowledge the support of the Austrian Science Fund (FWF) under research grant P 30878-N32. R. Quirchmayr acknowledges the support of FWF under research grant J 4339-N32. 6 C. I. MARTIN AND R. QUIRCHMAYR

References [1] A. Constantin, On the modelling of equatorial waves, Geophys. Res. Lett. 39, L05602 (2012). [2] A. Constantin, An exact solution for equatorially trapped waves, J. Geophys. Res. Oceans 117, C05029 (2012). [3] A. Constantin, Some three-dimensional nonlinear equatorial flows, J. Phys. Oceanogr. 43, 165– 175 (2013). [4] A. Constantin, Some nonlinear, equatorially trapped, nonhydrostatic internal geophysical waves, J. Phys. Oceanogr. 44(2), 781–789 (2014). [5] A. Constantin and R. S. Johnson, The dynamics of waves interacting with the Equatorial Undercurrent, Geophys. Astrophys. Fluid Dyn., 109(4), 311–358 (2015). [6] A. Constantin and R. S. Johnson, An exact, steady, purely azimuthal equatorial flow with a , J. Phys. Oceanogr. 46(6), 1935–1945 (2016). [7] A. Constantin and R. S. Johnson, An exact, steady, purely azimuthal flow as a model for the Antarctic Circumpolar Current, J. Phys. Oceanogr. 46(12), 3585–3594 (2016). [8] A. Constantin and R. S. Johnson, A nonlinear, three-dimensional model for ocean flows, mo- tivated by some observations of the Pacific Equatorial Undercurrent and thermocline, Physics of Fluids 29, 056604 (2017). [9] G. Danabasoglu, J. C. McWilliams and P. R. Gent, The role of mesoscale tracer transport in the global ocean circulation, Science (264), 1123–1126 (1994). [10] M. F. Cronin and W. S. Kessler. Near-Surface Shear Flow in the Tropical Pacific Cold Tongue Front, J. Phys. Oceanogr. 39, 1200–1215 (2009). [11] R. Farneti, T. L. Delworth, A. Rosati, S. M. Griffies and F. Zeng, The role of mesoscale eddies in the rectification of the Southern Ocean response to climate change, J. Phys. Oceanogr. 40(7), 1539–1557 (2010). [12] Y. L. Firing, T. K. Chereskin, and M. R. Mazloff, Vertical structure and transport of the Antarctic Circumpolar Current in Drake Passage from direct velocity observations, J. Geophys. Res. 116, C08015 (2004). [13] S. V. Haziot and K. Marynets, Applying the stereographic projection to modeling of the flow of the antarctic circumpolar current, Oceanography 31(3), 68–75 (2018). [14] S. V. Haziot, Study of an elliptic partial differential equation modelling the Antarctic circum- polar current, Discrete Contin. Dyn. Syst. 39(8), 4415–4427 (2019). [15] E. Howard, A. M. Hogg, S. Waterman and D. P. Marshall, The injection of zonal by buoyancy forcing in a Southern Ocean model, J. Phys. Oceanogr. 45, 259–271 (2015). [16] V. O. Ivchenko and K. J. Richards, The dynamics of the Antarctic Circumpolar Current, J. Phys. Oceanogr. 26, 753–774 (2012). [17] H.-C. Hsu and C. I. Martin, On the existence of solutions and the pressure function related to the Antarctic Circumpolar Current, Nonlinear Anal. 155, 285–293 (2017). [18] J. Klinck and W. D. Nowlin, Antarctic Circumpolar Current, Encyclopedia of Ocean Sciences, J. H. Steele, S. A. Thorpe and Karl K. Turekian Eds., Academic Press, 2001, 151–159. [19] D. P. Marshall, D. R. Munday, L. C. Allison, R. J. Hay and H. L. Johnson, Gill’s model of the Antarctic Circumpolar Current, revisited: The role of latitudinal variations in wind stress, Ocean Modelling 97, 37–51 (2016). [20] K. Marynets, The Antarctic Circumpolar Current as a shallow-water asymptotic solution of Euler’s equation in spherical coordinates, Deep- Res. Pt. II 160, 58–62 (2019). [21] J. Pedlosky, Geophysical , Springer, 1979. [22] R. Quirchmayr, A steady, purely azimuthal flow model for the Antarctic Circumpolar Current, Monatsh. Math. 187(3), 565–572 (2018). [23] H. Stommel, Wind drift near the , Deep Sea Res. 6, 298–302 (1960). [24] M. Tomczak and J. S. Godfrey. Regional Oceanography: an Introdution. Pergamon Press, Oxford, 1994. A STEADY STRATIFIED PURELY AZIMUTHAL FLOW REPRESENTING THE ACC 7

[25] A. F. Thompson, The atmospheric ocean: eddies and jets in the Antarctic Circumpolar Cur- rent, Phil. Trans. R. Soc. A 366, 4529–4541 (2008). [26] G. K. Vallis, Atmospheric and Oceanic Fluid Dynamics, Cambridge University Press, 2006. [27] W. B. White and R. G. Peterson, An Antarctic circumpolar wave in surface pressure, wind, temperature and sea-ice extent, Nature 380, 699–702 (1996). [28] J. O. Wolff, Modelling the Antarctic Circumpolar Current: eddy-dynamics and their parametrization, Environ. Modell. Software 14, 317–326 (1999).