1. Introduction
The aim or favorable characteristic of a simplified model should be to reproduce physical features known to be satisfied in more general and comprehensive models. Any buildup of a simplified model toward a “better” version should preferably not disrupt the physical properties that are satisfied in both general and simplified models. We say preferably because there might be some cases where a justification is at hand. This better model ends up halfway between the very simplified and general comprehensive models.
The conservation of energy and angular momentum are basic physical principles satisfied in comprehensive general models of nature. The shallow-water and reduced-gravity models in the f plane are simplified fluid dynamical models that reproduce these basic principles in its plain frictionless version, but the addition of lateral diffusion in the momentum equations might produce, depending on the parameterizations, inconsistencies. Shchepetkin and O’Brien (1996) analyze some consequences on the choice of the lateral frictional parameterization, and Gent (1993) shows a particular form that is consistent energetically, meaning that its inclusion in a closed domain, with either no-slip or free-slip boundary conditions, implies no energy increase for any state of the system. Schär and Smith (1993) deduce a form that is consistent with both principles. Warnings on the consequences over angular momentum have been profusely stated (e.g., Shchepetkin and O’Brien 1996; Wajsowicz 1993; Griffies 2004; Schär and Smith 1993). Here, we follow a parallel line of thought as Gent (1993) did for the energy budget but adding the principle of angular momentum conservation. The requirement for a reduced-gravity model to be consistent with the conservation of vertical angular momentum L is that its rate of change (i.e., dL/dt) should depend on the torques exerted by forces on the boundary and on the fluid interchange with the exterior for any volume in question: that is, exclusively on boundary conditions. Here, the meaning of boundary conditions is not necessarily those arising from some impositions on solid boundaries but as the conditions existing on the boundary of any subdomain under consideration. There is of course the possibility of including external forces: surface and bottom stresses and their corresponding torques. Here, we limit the scope to the unforced reduced-gravity model with lateral momentum diffusion and in the absence of stresses and mass exchange at the surface and bottom. Both external torques and fluid exchange must be accounted for via lateral boundary conditions; internal torques and internal redistribution of mass should not contribute to the variation of L. This requires a symmetric viscous stress tensor. It is also customary (Batchelor 1981) to set the frictional stress tensor with null trace, in which case the “pressure” or the isotropic contribution is independent of the inclusion or absence of frictional effects, but this is an overimposition when dealing with stress tensors arising from turbulence.
In Table 1, we show six explicit parameterizations of the lateral diffusion of momentum (cases I–V and VII), out of which three (III, IV, and VII) are consistent with the conservation of angular momentum and imply, as stated by Gent (1993), energy dissipation. Notice that cases I–V collapse into the same form when the layer thickness h is assumed uniform. Case VI is a family of symmetric stress tensor parameterizations that show the difficulty of fulfilling the energy requirement. The functions A and B in parameterization VI are arbitrary scalar functions of any combination of local variables (i.e., arbitrary scalar fields). Cases III, IV, and V are particular versions of case VI. Cases II and V are close relatives; any second-rank tensor, as stress tensors are, can be uniquely decomposed in the sum of its isotropic, symmetric, and antisymmetric components, and case V corresponds to the symmetric component of case II.
A set of parameterizations of lateral friction for reduce gravity models. The frictional stress tensor shown in the fourth column applies to the notation of Eq. (1). The last two columns indicate if the parameterization is consistent (OK), inconsistent (No), or indeterminate (?) with respect to the dissipation of energy and conservation of angular momentum.
Gent (1993) used the same parameterization as Gustafsson and Sundström (1978) and Bleck and Boudra (1981). Although he did not show explicitly the stress tensor, he recognized that the parameterization was the divergence of a stress tensor and decided in its favor instead of that used by Schär and Smith (1993) (case IV). In this reference, a traceless stress tensor is presented explicitly and shown to fulfill the requirements as stated here.
The 2D nondivergent flow model is slightly simpler than the reduced-gravity model. In the studies of Molenaar et al. (2004), van Heijst et al. (2003, 2006), and Clercx and van Heijst (2009) the conservation of angular momentum, as shown in those studies and stated here (i.e., depending exclusively on boundary conditions), is granted. In those studies, the parameterization in use is that of cases I–V because all of them collapse. The use of 2D nondivergent flow models avoid from the very start thickness variations in space and time, which is an ingredient that makes our analysis relevant.
Formal derivations via scaling assumptions, the law of the wall and the vertical integration of the Navier–Stokes equations, as shown by Gerbeau and Perthame (2001), Marche (2007), and Lucas and Rousseau (2008), show forms in favor of case V. This is the situation when removing bottom stress and several other terms included in their derivations: for example, additional isotropic contributions besides pressure in the stress tensor. However, in those derivations bottom stress is the primary frictional effect from which a secondary correction appears in the form of lateral diffusion of momentum. The purpose of this study is a warning call in the direction of the title, considering particularly that eddy parameterizations of turbulence are still open to a wide set of tests.
In section 2, we set definitions and the governing equations of the reduced-gravity model with the inclusion of eddy-induced transports. Section 3 shows the consequences on energy and angular momentum of the parameterizations defined in Table 1. In section 4, we discuss the results, and the last section gives the conclusions.
2. Governing equations and definitions
In the simpler situation of fully 2D nondivergent flows, as in examples by van Heijst et al. (2006), h′ ≡ 0,
A frequently used parameterization, which Gent (1993) proved to be inconsistent with energy requirements, replaces ∇ · s1 and ∇ · s2 in (1) by ϑh∇2u and ϑh∇2υ, respectively. As shown shortly, it fails the angular momentum requirement.
The use of the origin as the reference for the arm to define the angular momentum can be relaxed to any other reference point as long as it does not change with time. In this case, r is the position vector minus rO, where rO is the position of the reference point. The use of time-varying reference points adds the terms
If the boundary is static (i.e., vC · n = 0) and coincides with an impermeable boundary, there is no exchange of fluid with the exterior of the domain, v · n = 0 on ∂D, and several terms in (4) drop out. Likewise, if the subdomain of interest is a material blob (i.e., vC · n = v · n), several terms cancel.
In the absence of viscous effects (i.e., with s1 = s2 = 0) and with solid boundaries on the area under consideration dE/dt = 0 and
When vC · n = 0, interpretations of terms like
In the simpler 2D problem (i.e., nondivergent flow with constant and uniform h), a formal relationship between the angular momentum and integrals of vorticity forms is possible. This follows from the vorticity and boundary conditions uniquely setting the velocity field [see, e.g., Eqs. (9) and (10) in Molenaar et al. (2004) or (22)–(24) in Clercx and van Heijst (2009)]. The variable thickness or nondivergence of the flow prevents these reductions, and, to our knowledge, it is not possible in general to derive the angular momentum out of integrals of potential vorticity forms. Salmon (1998) shows how the conservation of potential vorticity is intimately related to a differential form of angular momentum, but with the arm defining the angular momentum in the same position as the column in consideration. The same is true for vorticity in general 3D flows, but again with the arm in the center of mass (Chatwinal 1973).
3. A set of parameterizations
The parameterizations listed in Table 1 show the relationship for several values of Θ and, with the exception of the first one, the corresponding stress tensor. Of these parameterizations, only the first one (I) does not allow to write it as the divergence of a tensor, and only parameterization II does not have a corresponding symmetric stress tensor. Forms I and II fail the principle of angular momentum conservation. Parameterizations III–VII follow from the use of a symmetric stress tensor and therefore are consistent with the principle of angular momentum conservation.
For the biharmonic parameterization, described in case VII of Table 1, the additional flux of energy that arises from the viscous terms is not only us1 + υs2. In this case, u∇ · s1 + υ∇ · s2 = ∇ · Fυ − d2, where Fυ = us1 + υs2 + ϑ∇[(ux − υy)2 + (υx + uy)2]/2,
4. Discussion
Setting A = 1 and B = 1 in case VI corresponds to case IV, which is the form used by Schär and Smith (1993) and satisfies Eq. (8) automatically {i.e., [(ux − υy)2 + (υx + uy)2] ≥ 0}. Setting A = h and B = 1 is case V, and it also satisfies Eq. (8) [see Eqs. (11)] because h is a nonnegative function. Although Gent (1993) does not explicitly exhibits a stress tensor, the corresponding of case II implies the same parameterization. A tensor of this kind can be uniquely decomposed as the sum of an isotropic tensor, a symmetric tensor, and an antisymmetric tensor. The stress tensor of case V is the symmetric component of case II. Case III corresponds to the choice A = 1 and B = h.
Parameterizations IV and V are consistent with the principles of energy and angular momentum conservation, but case V has an awkward form in the momentum equations viscous terms; one component depends on the cross product of the thickness and the orthogonal velocity component gradients (see Table 1). This awkward dependency exists for all nontrivial choices of the function A. Only with a trivial choice of function A do the viscous terms in the momentum equations have a familiar form. Only with a trivial choice of function B is the energy requirement directly guaranteed.
Case VII is peculiar. It shows a viscous-dependent energy flux that has more terms than the immediately recognized contribution us1 + υs2.
The use of a stress tensor that is a linear combination of different cases is also possible. If the individual cases involved in the linear combination satisfy the decomposition
The neglect of bottom stress is an assumption of the reduced-gravity model at hand, an assumption that seems reasonable because the slippage of the active layer is over heavier fluid and not over a riverbed as in Gerbeau and Perthame (2001), Marche (2007), and Lucas and Rousseau (2008). These derivations of viscous shallow-water equations via integration of the Navier–Stokes equations show the lateral viscous terms as a secondary corrections, using the law of the wall, of the primary frictional effect: that due to bottom stress. Thus, in the context of an active layer slipping without friction over an inactive heavier layer, the form of the lateral frictional stress cannot be formally supported by those studies. Moreover, if friction is allowed between the active and inactive layers the reduced-gravity model at hand is incomplete.
The stress with the underlying fluid when the active layer is very thin or forms an edge can hardly be neglected. Therefore, under such conditions, this model becomes unrealistic (i.e., of questionable usefulness). Even more, insisting within this model in the limit of null thickness, the
4. Conclusions
As shown by many authors, the consistency with the principle of angular momentum conservation implies restrictions on the allowed frictional parameterizations that are easy to fulfill (i.e., a symmetric stress tensor). Here, we show that the freedom to choose the stress tensor is limited, given the restriction coming from energy considerations. A broad family of tensors, which depends on two arbitrary weighting functions, as depicted by case VI in Table 1, shows that, to assure consistency with the energy dissipation, (i) function B must be a trivial constant function; (ii) function A is quite free, with the only requirement of definite sign; but (iii) any nontrivial function A imposes a rather awkward friction form in the momentum equations. Even very simple choices of a function A, as the resulting form (A = h) in the symmetric component of the stress tensor proposed by Gent (1993), produces an unusual frictional term (see case V).
The very common parameterization of friction, case I, fails to requirements regarding energy dissipation (Gent 1993) and angular momentum. The close relative to case I that is consistent with energy dissipation as shown by Gent (1993) is case II, but it is not consistent with the principle of angular momentum. Case II can easily be transformed to case V, which is consistent with both energy and angular momentum but at the expense of rather awkward viscous terms in the momentum equations. The parameterization used by Schär and Smith (1993), which in case IV is simple and fulfils the requirements with energy and angular momentum.
The biharmonic parameterization (case VII) is consistent with the angular momentum conservation principle and with energy dissipation. The parameterization of the eddy-induced transport within this reduced-gravity model is unnecessary, because the actual variables to be solved are the full transports, without solving separate contributions.
Proper parameterization of Reynolds stresses in ocean models remains an unsolved issue (Griffies 2004; Marshall and Adcroft 2010; Eden and Greatbatch 2008). Our purpose here has been simply to draw attention to the properties of some commonly used horizontal viscosity parameterizations employed in layer models, and we have not dwelled into the physical basis that relate eddy stresses with the formulas given in Table 1. Users of these simple models should favor the use of formulas that preserve basic physical principles such as the ones discussed here.
Acknowledgments
Funded by CONACYT Grant U50204-F. Encouragement and comments by Gert-Jan van Heijst are greatly appreciated. The author gives thanks to all of the comments from the anonymous reviewers, they helped improve the paper.
REFERENCES
Adcock, S. T., and D. P. Marshall, 2000: Interactions between geostrophic eddies and the mean circulation over large-scale bottom topography. J. Phys. Oceanogr., 30, 3223–3238.
Batchelor, P. R., 1981: An Introduction to Fluid Mechanics. Cambridge University Press, 615 pp.
Bleck, R., and D. B. Boudra, 1981: Initial testing of a numerical ocean circulation model using a hybrid (quasi-isopycnic) vertical coordinate. J. Phys. Oceanogr., 11, 755–770.
Chatwinal, P. C., 1973: The vorticity equation as an angular momentum equation. Math. Proc. Cambridge Philos. Soc., 74, 365–367.
Clercx, H. J. H., and G. J. F. van Heijst, 2009: Two-dimensional Navier–Stokes turbulence in bounded domains. Appl. Mech. Rev., 62, 020802, doi:10.1115/1.3077489.
Eden, C., and R. J. Greatbatch, 2008: Towards a mesoscale eddy closure. Ocean Modell., 20, 223–239.
Gent, P. R., 1993: The energetically consistent shallow-water equations. J. Atmos. Sci., 50, 1323–1325.
Gerbeau, J.-F., and B. Perthame, 2001: Derivation of viscous Saint-Venant system for laminar shallow water; Numerical validation. Discrete Contin. Dyn. Syst., 1B, 89–102.
Griffies, S. M., 2004: Fundamentals of Ocean Climate Models. Princeton University Press, 518 pp.
Gustafsson, B., and A. Sundström, 1978: Incompletely parabolic problems in fluid dynamics. SIAM J. Appl. Math., 35, 343–357.
Jiménez Domínguez, A., 2008: Momento angular en un modelo de gravedad reducida con fricción lateral. M.S. thesis, CICESE, 34 pp.
Lucas, C., and A. Rousseau, 2008: New developments and cosine effect in the viscous shallow water and quasi-geostrophic equations. Multiscale Model. Simul., 7, 793–813.
Marche, F., 2007: Derivation of a new two-dimensional viscous shallow water model with varying topography, bottom friction and capillary effects. Eur. J. Mech., 26B, 49–63.
Marshall, D. P., and A. J. Adcroft, 2010: Parameterization of ocean eddies: Potential vorticity mixing, energetics and Arnold’s first stability theorem. Ocean Modell., 32, 188–204.
Molenaar, D., H. J. H. Clercx, and G. J. F. van Heijst, 2004: Angular momentum of forced 2D turbulence in a square no-slip domain. Physica D, 196, 329–340.
Salmon, R., 1998: Lectures in Geophysical Fluid Dynamics. Oxford University Press, 378 pp.
Schär, C., and R. B. Smith, 1993: Shallow-water flow past isolated topography. Part I: Vorticity production and wake formation. J. Atmos. Sci., 50, 1373–1400.
Shchepetkin, A. F., and J. J. O’Brien, 1996: A physically consistent formulation of lateral friction in shallow-water equation ocean models. Mon. Wea. Rev., 124, 1285–1300.
van Heijst, G. J. F., H. J. H. Clercx, and S. R. Maassen, 2003: A note on the effects of solid boundaries on confined decaying 2D turbulence. Nonlinear Processes in Geophysical Fluids, O. U. Velasco Fuentes, J. Sheinbaum and J. Ochoa, Eds., Kluwer Academic, 305-324.
van Heijst, G. J. F., H. J. H. Clercx, and D. Molenaar, 2006: The effects of solid boundaries on confined two-dimensional turbulence. J. Fluid Mech., 554, 411–431.
Wajsowicz, R. C., 1993: A consistent formulation of the anisotropic stress tensor for use in models of the large-scale ocean circulation. J. Comput. Phys., 105, 333–338.