1. Introduction
Meridional heat transport due to geostrophically balanced, swirling motions relative to mean flows contributes significantly to the oceanic and atmospheric energy balance (Wunsch and Ferrari 2018; Held 2019). In oceanographic context, these swirling flows are referred to as mesoscale eddies. When viewing these eddies as macroturbulence (e.g., Rhines 1977; Salmon 1980; Held 1999), one may represent their transport by the turbulent diffusion formulation, with an eddy diffusivity encapsulating the bulk effects of eddy stirring that ultimately lead to net transport [see, e.g., Jansen et al. (2015) for a recent synthesis]. To constrain the turbulent transport, an important task is then to understand how the diffusivity (or equivalently the eddy velocity and length scales via the mixing length theory) is related to mean flows and environmental conditions (Vallis 2021).
To develop understanding for the eddy diffusivity, a reduced-physics system, known as homogeneous two-layer quasigeostrophic (2LQG) turbulence, has been used as a testbed. The 2LQG turbulence aims to represent a patch of eddy field in which spatial and temporal variations of the mean flow are sufficiently small such that eddy statistics may be considered to be locally homogeneous. It is thus justified to break the full inhomogeneous problem into piecewise uniform sections where one focuses on relating the eddy properties with an externally imposed, locally uniform mean state. There is supporting evidence for the applications of this local approach to global inhomogeneous flows (e.g., Pavan and Held 1996; Chang and Held 2019; Gallet and Ferrari 2021).
For 2LQG turbulence, there is an extensive literature on the scaling theories of eddy diffusivity in two asymptotic regimes. From a perspective of turbulent cascade, these two regimes correspond to different stopping mechanisms of a barotropic inverse cascade: by damping due to bottom drag (e.g., Larichev and Held 1995; Held 1999; Arbic and Flierl 2004; Chang and Held 2019; Chen 2023) and by the β effects that channel energy into Rossby waves/zonal modes (e.g., Vallis and Maltrud 1993; Held and Larichev 1996; Lapeyre and Held 2003). In this study, these two regimes are referred to as drag-controlled and β-controlled, respectively. The general idea is that the diffusivity is set by the halting scale because it represents the energy-containing eddies. Theories were then built upon balance constraints between energy production, dissipation, and cascade [e.g., motivated by Salmon’s (1980) energy pathways], with these quantities expressed in terms of the halting scale and associated barotropic energy level. Much of the theoretical development has focused on the limits where one of the cascade-halting mechanisms dominates. Note that there were also recent advances in a different perspective where the diffusivity was predicted based on interactions between pairs of dilute vortices (referred to as the vortex-gas model; Gallet and Ferrari 2020, 2021; Thompson and Young 2006). In this study, we take on the cascade viewpoint as a followup of Chen (2023), but comparisons with the vortex-gas model will be made.
The theories targeting asymptotic scaling behaviors clearly will encounter problems when applying to transitions where bottom drag and β influences are both at work. An apparent need for a more complete theory has motivated further developments. Recently, Chang and Held (2021, hereafter CH21) and Gallet and Ferrari (2021), hereafter GF21) both proposed theories for the diffusivity that have explicit drag and β dependencies. The central idea is motivated by the work of Ferrari and Nikurashin (2010): The diffusivity is bounded by the f-plane value and is suppressed by β, with the suppression strength governed by some dimensionless measures of Rossby wave phase speed or wave-turbulence crossover scale (see section 2c). The advances made by CH21 and GF21 have improved the overall representation of the eddy diffusivity. In both studies, the predicted diffusivities showed quantitative agreement with numerical simulations.
In comparison with the extensive works on diffusivity formulations, less attention has been given to understanding the quantitative properties of zonal jets that are known to emerge from β-plane turbulence. In particular, the understanding for the relative strength of jets and eddies as well as their relative contributions to dissipation across the parameter space is still quite limited.
For instance, a hypothesis of jet dominance in energy dissipation was proposed to help defend a scaling prediction that eddy properties are drag independent. But the validity of this hypothesis is largely unverified. The drag-free prediction is derived from a theory, referred to as the β-scaling, put forth by Held and Larichev (1996, hereafter HL96). The β-scaling was formulated assuming the existence of a Kolmogorovian spectrum with a dissipation ε setting the rate at which energy flows through the system and that the eddy length scale is limited by Rossby wave-turbulence crossover phenomenology (see section 2b). It gives analytical expressions for barotropic eddy velocity, mixing length, and diffusivity in terms of ε and β and has shown predictive skills when applied to a variety of turbulent flows (e.g., Barry et al. 2002; Lapeyre and Held 2003; Zurita-Gotor 2007; Chang and Held 2022). However, HL96 demonstrated that, when the scaling is combined with energy balance, the explicit dependency on ε can be eliminated. The resulting dimensionless eddy scales become only a function of β [see Eqs. (3) and (4) in section 2b]. The complete lack of drag sensitivity in the scaling is surprising and has been a subject of further examinations (e.g., Thompson and Young 2007). It is surprising because true termination of inverse cascade ultimately requires energy dissipation. But the dissipative effects are not explicit in the scaling formulations. A defense for the drag insensitivity is to hypothesize that a majority of energy dissipation occurs in zonal jets, thereby shielding eddies from a direct damping effect of bottom drag (HL96; see section 2b for mathematical expressions). This argument is referred to as the jet dominance hypothesis in this work. Note that, while the hypothesis is conceptually appealing, its validity has not been carefully examined.
For zonal jet energetics, it is well established that the standard QG theory of eddy–mean-flow interactions provides a useful framework to diagnose the generation of zonal flows from a background eddy field. From this framework, a central result is that the bottom drag exerted on a zonal mean flow is balanced by the vertical integral of eddy potential vorticity (PV) fluxes (e.g., see Vallis 2017, ch. 15.2–15.4; Galperin and Read 2019). Here the eddy PV fluxes represent the divergence of eddy momentum fluxes in the horizontal and form stresses in the vertical. The vertical integral therefore subsumes the momentum transfer due to eddies over a column that could force zonal jets [e.g., a classic example of this is the atmospheric eddy-driven jet; see again Vallis (2017)]. However, an estimate of jet velocity needs knowledge of eddy PV fluxes, which are generally not known a prior. This complicates the use of the momentum balance constraint to interpret the jet regime’s parameter sensitivity. On the other hand, a global (domain integrated) energy balance could provide an integral constraint to be exploited. If quantitative formulations for diffusivity and eddy velocity are available, for example from the β-scaling, and jet/eddy dissipation can be separated, it seems possible to infer jet speed from the global energy balance. In this study we will explore such a possibility, aiming to better understand jet’s parameter sensitivity.
Here we set out to develop a simple representation of thermal diffusivity in 2LQG turbulence on a β plane. The goals are to use the diffusivity formulation to 1) quantify the turbulent flow regimes, 2) develop a theory for zonal jet speed from energy balance, and 3) use the velocity scale estimates to better understand jet/eddy contributions to energy dissipation. The numerical experiments described in GF21 are expanded. Guided by the experiments, an empirical combination of two asymptotic theories leads to a diffusivity formulation that is quite simple and robust. From the formulation, a diagram separating the drag- and β-controlled regimes is readily constructed. The regime diagram also helps identify the parameter range over which the β scaling for eddy scales is applicable. We aim to show that the quantification of flow regimes, diffusivity, and eddy velocity allows a reasonably skillful estimate of jet speed to be developed and interpreted. It permits evaluations of the jet dominance hypothesis, from which a modest revision is proposed. The relative contributions of jets and eddies to energy dissipation across the parameter space are also explored. The rest of this paper is organized as follows. In section 2, we briefly review the asymptotic as well as modern diffusivity theories to be used and compared against. In section 3, numerical experiments, diagnostic metrics, and model validations are described. Section 4 presents a simple diffusivity formulation. In section 5, the formulation is then used to tackle the tasks of quantifying turbulent flow regimes and obtaining theoretical estimates for eddy and jet velocities in β-controlled turbulence. The roles of jets and eddies in energy dissipation are also examined. Last, in section 6, mapping of a subspace referred to as the zonostrophic regime is presented, along with a brief discussion of the implications for eddy parameterization.
2. A summary of theories to be applied and evaluated
a. Drag-controlled (f-plane) regime
b. β-controlled regime (β scaling)
An important feature shared by Eqs. (3) and (4) is the lack of an explicit drag dependency. If the dissipation in energy balance depends only on drag coefficient and eddy velocity (i.e.,
c. Combined effects
As introduced in section 1, theories recently proposed by CH21 and GF21 considered the thermal diffusivity with a general drag and β dependencies. These theories represent the state of knowledge. Predictions from them will be used as benchmarks for comparisons (see section 4). Below we summarize the relevant formulations and parameters. Detailed derivations shall refer to the original papers.
To obtain a prediction, one first determines the scaling coefficient for
To obtain a prediction, one first determines the scaling coefficients a0 and a1 for the f-plane mixing length and diffusivity and then apply Eq. (6). GF21 used a0 = 2.5 and a1 = 2. They reported that the empirical suppression function F(B) in Eq. (6) with a2 = 2.5 captures the behavior of
3. Methods
a. Numerical experiments
Design of the numerical experiments follows Chen (2023) but is extended to consider the effects of β suppression. The evolution of 2LQG turbulence is governed by the standard two-layer quasigeostrophic potential vorticity (QGPV) equations on a β plane. We consider a basic state of two equal-thickness layers imposed with constant zonal mean shear of ±U. Dissipative terms identical to those in GF21 and Chen (2023) are included to the evolution equations: A bottom drag term is confined to the lower layer, which parameterizes the Ekman spindown to allow eddy fields to equilibrate. For numerical stability, hyperviscous damping is added for both layers to remove enstrophy accumulation at grid scales [see GF21’s Eqs. (1)–(4)]. Simulations are carried out in a 2πL × 2πL doubly periodic box. Note that, as in Chen (2023), both quadratic and linear drag forms are considered, but the focus is on the former.
The governing equations are made dimensionless using U and L and are integrated forward in time using a well-validated spectral solver Dedalus (Burns et al. 2020). The dimensionless equations are identical to those in Chen (2023) (see his appendix A), except that the inclusion of β modifies the background PV gradients. In the dimensionless form, the upper- and lower-layer PV gradients are
b. Metrics and model validation
4. Regime transition and a diffusivity formulation
a. Drag-dependent regime transition
We first use the experiments to illustrate that the diffusivity is approximately constrained by the drag- and β-controlled asymptotes. In Fig. 1a, the responses of diffusivity to varied
(a) Responses of dimensionless thermal diffusivity
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
A different way to evaluate the bound provided by Lapeyre and Held’s (2003) drag-free scaling is to plot the ratio of
Note however that the drag sensitivity does not stay minimal as
Given that the diffusivity for a given
Structure of baroclinic streamfunction (i.e., proportional to temperature) at equilibrium for (a),(b)
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
However, the value of critical
The main points to be made here are the following. First, for a given
b. A simple formulation for eddy diffusivity applicable to 2LQG and 2D turbulence
The diffusivity representation of Eq. (10) is tested against numerical experiments. Despite its simplicity, the agreement over the entire range of Dτ is very good. In Fig. 3a, for a given
Evaluations of the (a),(b) diffusivity formulation of Eq. (10) and (c),(d) GF21’s theory of Eq. (6). (left) The
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
For comparisons, the same numerical dataset is used to evaluate the predictions of GF21’s and CH21’s theories [Eqs. (5) and (6)], with the statistics summarized in Table 1. The comparisons with GF21 are shown in Figs. 3c and 3d. As described in section 2c, the coefficients in Eq. (6) are obtained from best fits. It can be seen in Fig. 3 that GF21’s predictions are in agreement with the data, but the comparison is not as tight as the predictions using Eq. (10). The points are clearly more spread out from the diagonal line in Fig. 3d than in Fig. 3b.
The comparisons are made more quantitative by evaluating the root-mean-square error (RMSE). The RMSEs for
A summary of root-mean-square error (RMSE) among various predictions of
The same representation of Dτ in Eq. (10) is applicable to cases with linear drag and to forced-dissipative two-dimensional (2D) turbulence. For linear drag and 2D turbulence, we compare the predictions using Eq. (10) with GF21’s and KJ17’s theories, respectively. Details are described in appendix B. Here we only report the RMSEs to summarize the comparisons: For linear drag, the RMSE is generally comparable to GF21’s theory, with a value of 0.120 versus 0.103 for the best-fit coefficients. For 2D turbulence, the RMSE is also comparable to KJ17’s, having small RMSEs of 0.065 and 0.068, respectively.
c. Physical interpretations
The above comparisons lead us to conclude that the simple diffusivity representation in Eq. (10), albeit empirical, is quite robust and reasonably accurate. It is robust because it is constrained by two well-established asymptotes. The robustness is further supported by the fact that the same functional form is applicable to 2LQG for both drag forms and to 2D turbulence. The empiricism involved in Eq. (10) is far from satisfactory, but some degree of empiricism is not uncommon in diffusivity formulations [e.g., GF21; CH21; see comments by Vallis (2021)]. A heuristic argument for how the empirical form may arise will be discussed in section 7.
Given that
5. Utility of the diffusivity formulation
As stated in the introduction, the main goals of this study are to use the diffusivity formulation to help 1) quantify the turbulent flow regimes, 2) develop a global estimate of zonal jet speed, and 3) better understand jet/eddy contributions to energy dissipation. With the
a. A regime diagram and eddy scales
From Eq. (10) and Fig. 1, we have seen that the transition between drag- and β-controlled asymptotes occurs at
A diagram to classify the turbulent flow regimes in
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
The regime diagram in Fig. 4 has identified the parameter range over which β effects exert a stronger control over frictional processes in setting the eddy diffusivity (i.e., to the right of
Test of the β scaling of Eq. (2) in the β-controlled regime as identified using the regime diagram of Fig. 4. In both panels, the experiments in the β-controlled regime (i.e.,
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
b. A jet speed theory from energy balance
In Eq. (14a) we have assumed zonal jets to be approximately barotropic. Analyses suggest that this assumption is a reasonable starting point. We calculate the ratio of rms lower-layer and barotropic zonal velocities (
c. Testing the jet speed estimates
In Fig. 6, the jet speed predictions in Eqs. (14) and (16) are tested against the simulations. In Fig. 6a, there is reasonable agreement between the simulated and predicted
Evaluations of an energy-balance-based theory for zonal jet speed. (a) Predictions of rms barotropic jet speed from Eq. (14) are compared with the simulation results. The cases are all in the β-controlled regime (
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
To illustrate the parameter dependency, next we show examples of how
In contrast to the jet’s clear negative response to drag variations, the eddy velocity stays largely unchanged (Fig. 6c). The drag-free scaling of Eq. (15) represents the eddy velocity well (dashed line). Inserting
Physically, the different drag sensitivity of jets and eddies may be understood as follows. For cases deep into the β-controlled regime (
When
There is a subtle point to note here. The largely negative dependency of
d. Jet versus eddy contributions to energy dissipation: A revision of jet dominance hypothesis
In Figs. 6c and 6b, we have seen hints for the departure of HL96’s jet dominance hypothesis. In Fig. 6c where all cases satisfy
To examine the jet dissipation fraction when eddy properties are drag insensitive, we may restrict our attention to
Evaluations of (a) jet fraction of energy dissipation and drag-independent properties in (b) eddy diffusivity and (c) eddy velocity. In these panels, εjet/ε,
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
However, among these cases, energy dissipation is often not dominated by zonal jets. In Fig. 7a, the jet fraction εjet/ε can be well below 0.5. It varies between 0.07 and 0.68, with a mean of 0.40 ± 0.18. The calculations suggest that eddies play a nonnegligible role in energy dissipation, inconsistent with HL96’s jet dominance hypothesis. Note that the above conclusion does not change if we include
To sum up, the analyses in Fig. 7 and in section 5c suggest a modest modification of HL96’s jet dominance hypothesis. For cases with
6. Discussion
a. A quantification of the zonostrophic regime
The analyses in section 5d have suggested that jet dominance in dissipation is not a requirement to maintain drag-insensitive eddy statistics. Yet, a regime of jet dominance has attracted considerable attention in the literature, for its highly anisotropic flow fields and energy spectrum. Such a regime, referred to as zonostrophic, was typically identified using a length scale ratio for 2D turbulence (e.g., Galperin et al. 2006; Sukoriansky et al. 2007; Scott and Dritschel 2012). Here we will use jet–eddy dissipation ratio as a direct quantification for 2LQG flows. A connection between dissipation and length scale ratios will be made.
Using the jet–eddy partitioning described in section 5d, we map out the dissipation ratio εjet/εe in the
(a) Quantification of jet–eddy dissipation ratio, εjet/εe, in the
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
The jet theory developed in section 5b could be used to estimate the regime transition and help interpret the parameter dependency. By Eqs. (12) and and (13), εjet and εe are estimated as
From the rough scaling, the parameter dependency of the zonostrophic regime could be interpreted as follows. For cases in the β-controlled regime satisfying
It is interesting to note that the scaling criterion of Eq. (17) could be converted to a length scale ratio established to identify 2D zonostrophic turbulence. In 2D flows, zonostrophic has been characterized by a highly anisotropic kinetic energy (KE) spectrum, with a spectral peak of zonal flow KE surpassing that of residual eddies (Galperin et al. 2006; Sukoriansky et al. 2007). If the zonal flow and eddy KE spectrum take a k−5 and k−5/3 slope, respectively [see Eq. (13.21) in a review by Galperin and Read (2019)], it is easy to verify that the two spectra intersect at a wavenumber kβ [
b. Implications for the eddy parameterization problem
The scaling estimate for εjet/εe in Eq. (17) and shown in Fig. 8b may have implications for oceanic eddy parameterizations. From the perspectives of climate simulations, an adequate representation of oceanic eddy fluxes, typically cast in a form of eddy diffusivity times mean gradients, is critically needed. This is because state-of-the-art climate models that target long-term simulations cannot resolve those eddies (e.g., see Jansen et al. 2015). Like the approaches described in section 2, modern eddy parameterizations often used energy balance to constrain the diffusivity (e.g., Cessi 2008; Marshall and Adcroft 2010; Jansen et al. 2015; Mak et al. 2018). A common ingredient is to express the energy dissipation as a function of eddy kinetic energy itself (e.g., ∼μE3/2 or κE2 where E is barotropic eddy energy) or to use a linear formulation with an unknown damping rate that subsumes all dissipative processes (e.g., Mak et al. 2022). However, if the results of 2LQG turbulence carry over, the unresolved fluctuations will include spontaneously generated zonal jets, especially under conditions of weak drag and large β (i.e., toward a zonostrophic regime). This implies that expressing dissipation in terms of eddy energy alone may overlook energy sinks through zonal jets.
The scaling in Eq. (17) may offer a crude way to account for the overlooked energy sinks. For example, with Eq. (17), the total dissipation including jet and eddy contributions may be expressed in terms of εe as
7. Summary and outlook
This study uses a simple diffusivity formulation to examine flow regime transition and eddy–jet energy partitioning in homogeneous 2LQG turbulence on a β plane. An estimate of the root-mean-square zonal jet speed is developed from a global energy balance.
The formulation for thermal diffusivity is a combination of two existing asymptotic theories, empirically constructed to satisfy the constraints as suggested by numerical experiments. In the limit of vanishing β (referred to as the drag-controlled regime), the diffusivity approaches an upper bound of a f-plane theory
The simple formulation enables a straightforward quantification of flow regimes, whereby the parameter range appropriate for an existing eddy scaling is readily identified. Via Eq. (10), the dimensionless diffusivity is mapped out in
With the quantitative representations for diffusivity and eddy velocity in hand, the rms speed of zonal jets emerging from β-controlled turbulence can be inferred from a global energy balance [Eqs. (14) and (16)]. The theory is able to provide a reasonable leading-order estimate for the jet speed in the simulations (Fig. 6) and clarify the jet’s parameter dependency: the jet speed scales inversely with drag strength and β. As
Building from the foregoing theoretical considerations, we proceed to investigate two questions concerning the jet–eddy energy partitioning in β-controlled turbulence. One is about whether jet dominance in dissipation is a requirement to keep eddy statistics drag-free (i.e., the jet dominance hypothesis; in section 5d). The other is to quantify and estimate the parameter space in which zonal flows are really energetically dominant relative to eddies—a subspace referred to as the zonostrophic regime (in section 6a).
The jet dominance hypothesis is first evaluated. Held and Larichev (1996) argued that, if dissipation depends only on eddy velocity as typically formulated, then the drag-free property of diffusivity and eddy velocity derived from the β scaling [Eqs. (3) and (4)] cannot be simultaneously satisfied in energy balance. This has led to a hypothesis that dissipation must occur in zonal jets, not eddies. It is shown here that, using a criterion of
We are also able to characterize the zonostrophic regime. This regime has been described for its highly anisotropic KE spectrum and identified by a length scale ratio for 2D flows, but these basic properties have not been fully examined for 2LQG turbulence. Again in the
There are a number of unresolved issues that are worth pointing out here. An obvious one is the empiricism involved in the diffusivity formulation. Equation (10) was not derived from first principles but empirically constructed to satisfy the asymptotic behavior seen in the simulations. That is, as
Another open question is the link between the energy-based estimate of jet speed and QG eddy-mean interaction theory for zonal mean flows. In section 5b, we have developed a theory for the rms jet speed from the global energy balance [see Eqs. (12)–(14)]. There is a well-established local constraint that is not used in this study. Here, global and local refer to domain average and zonal mean properties, respectively. The local constraint comes from momentum balance of zonal mean flows in the standard QG eddy–mean interaction framework. Specialized to two-layer QG flows, the requirement of no net meridional transport gives a constraint of
Last, we should also point out that the drag-free diffusivity scaling proposed by Lapeyre and Held (2003) [Eq. (4)] provides a better asymptotic bound than that by HL96 [Eq. (3)]. In Fig. 1a, there is a steeper drop in diffusivity toward larger
Acknowledgments.
Two anonymous reviewers provided constructive comments that improved this work substantially. Discussions with CY Chang (GFDL) and Kaushik Srinivasan (UCLA) about an early version of this work were helpful. This work is supported by the National Science and Technology Council of Taiwan through Grants MOST 108-2611-M-002-022-MY4 and NSTC 112-2611-M-002-015.
Data availability statement.
The source code for solving the QGPV equations by Dedalus and simulation outputs are available upon request from the author.
APPENDIX A
An Approximate Solution of Chen’s (2023) f-Plane Theory for Eddy Diffusivity
Here we briefly describe an approximate solution of Chen’s (2023) f-plane theory for eddy diffusivity in 2LQG turbulence. In what follows, we will develop approximations, test the solution against numerical experiments under a condition of vanishing β, and discuss its linkages with prior studies.
In Eq. (A1), the first equation is a production-dissipation balance, with a correction of partial barotropization incorporated into the dissipation estimates (i.e., −2VfU term, allowing eddies to be baroclinic; see Chen 2023) and μ being a quadratic drag coefficient. The second describes a cascade-production balance, modified to empirically allow the cascade fraction εc/εp to increase with cascade extent. In Eq. (A1b), the dimensionless drag strength
Evaluations of an approximate solution for Chen’s (2023) f-plane theory. The experimentally derived (a) barotropic eddy velocity, (b) mixing length, and (c) diffusivity are plotted against the drag strength
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
The goodness of the approximate solution is confirmed in Fig. A1. To obtain predictions from Eqs. (A3) and (A4), we use coefficients identical to Chen (2023), with c0 = 0.31, c2 = 2, and m = 1/7. The numerical experiments with
It is worth pointing out that the above theory can recover the scaling results of prior studies. When assuming complete eddy barotropization and existence of an inertial range, we neglect the dissipation correction [i.e., +2 term in Eq. (A4)] and set m = 0. The above theory is reduced to Held’s (1999) scaling, with
APPENDIX B
Applicability to Linear Drag and Two-Dimensional Turbulence
The analyses below aim to show that the same diffusivity (Dτ) formulation in Eq. (10) is applicable to cases with linear bottom drag and to forced-dissipative two-dimensional (2D) turbulence. For 2LQG turbulence with linear drag, we have carried out new experiments that are similar to those in GF21 but expand the drag range to O(1), as described in section 3a. For 2D turbulence, numerical data are obtained by digitizing KJ17’s Fig. 3. Predictions of Eq. (10) are then compared with GF21’s and KJ17’s theories. The predictions are obtained as follows. In the case of linear drag,
Like for the quadratic drag shown in Fig. 3, the variations of
As in Fig. 3 in the main text, but evaluating the diffusivity formulation of Eq. (10) when applied to (a),(b) 2LQG turbulence with linear drag and to (c),(d) 2D turbulence with quadratic drag. In (a) and (b),
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0110.1
Again we can calculate the RMSE and compare it with GF21’s theory. Like for quadratic drag, we try two sets of scaling coefficients to obtain diffusivity predictions from GF21’s Eqs. (11) and (17): original values and values from best fits following the procedure of GF21 (see section 2c and Table 1’s caption). The RMSE using Eq. (10) is 0.120, equivalent to a mean drift by a factor of 1.32 (Table 1). This error is comparable to GF21’s RMSE of 0.103 and 0.175, with best-fit and original scaling coefficients, respectively. Overall, the comparisons suggest that, for linear bottom drag, the diffusivity formulation in Eq. (10) has predictive skills similar to GF21’s.
To represent tracer diffusivity in 2D turbulence, the formulation in Eq. (10) also shows skills comparable to the more elaborated theory of Kong and Jansen (2017). Figure B1c reproduces KJ17’s Fig. 3, where the tracer diffusivity normalized by the f-plane scaling is plotted against a controlling dimensionless number σ (
REFERENCES
Arbic, B. K., and G. R. Flierl, 2004: Baroclinically unstable geostrophic turbulence in the limits of strong and weak bottom Ekman friction: Application to midocean eddies. J. Phys. Oceanogr., 34, 2257–2273, https://doi.org/10.1175/1520-0485(2004)034<2257:BUGTIT>2.0.CO;2.
Barry, L., G. C. Craig, and J. Thuburn, 2002: Poleward heat transport by the atmospheric heat engine. Nature, 415, 774–777, https://doi.org/10.1038/415774a.
Burns, K. J., G. M. Vasil, J. S. Oishi, D. Lecoanet, and B. P. Brown, 2020: Dedalus: A flexible framework for numerical simulations with spectral methods. Phys. Rev. Res., 2, 023068, https://doi.org/10.1103/PhysRevResearch.2.023068.
Cessi, P., 2008: An energy-constrained parameterization of eddy buoyancy flux. J. Phys. Oceanogr., 38, 1807–1819, https://doi.org/10.1175/2007JPO3812.1.
Chang, C.-Y., and I. M. Held, 2019: The control of surface friction on the scales of baroclinic eddies in a homogeneous quasigeostrophic two-layer model. J. Atmos. Sci., 76, 1627–1643, https://doi.org/10.1175/JAS-D-18-0333.1.
Chang, C.-Y., and I. M. Held, 2021: The parameter dependence of eddy heat flux in a homogeneous quasigeostrophic two-layer model on a β plane with quadratic friction. J. Atmos. Sci., 78, 97–106, https://doi.org/10.1175/JAS-D-20-0145.1.
Chang, C.-Y., and I. M. Held, 2022: A scaling theory for the diffusivity of poleward eddy heat transport based on Rhines scaling and the global entropy budget. J. Atmos. Sci., 79, 1743–1758, https://doi.org/10.1175/JAS-D-21-0242.1.
Charney, J. G., 1971: Geostrophic turbulence. J. Atmos. Sci., 28, 1087–1095, https://doi.org/10.1175/1520-0469(1971)028<1087:GT>2.0.CO;2.
Chen, S.-N., 2023: Revisiting the baroclinic eddy scalings in two-layer, quasigeostrophic turbulence: Effects of partial barotropization. J. Phys. Oceanogr., 53, 891–913, https://doi.org/10.1175/JPO-D-22-0102.1.
Ferrari, R., and M. Nikurashin, 2010: Suppression of eddy diffusivity across jets in the Southern Ocean. J. Phys. Oceanogr., 40, 1501–1519, https://doi.org/10.1175/2010JPO4278.1.
Gallet, B., and R. Ferrari, 2020: The vortex gas scaling regime of baroclinic turbulence. Proc. Natl. Acad. Sci. USA, 117, 4491–4497, https://doi.org/10.1073/pnas.1916272117.
Gallet, B., and R. Ferrari, 2021: A quantitative scaling theory for meridional heat transport in planetary atmospheres and oceans. AGU Adv., 2, e2020AV000362, https://doi.org/10.1029/2020AV000362.
Galperin, B., and P. L. Read, 2019: Zonal Jets: Phenomenology, Genesis, and Physics. Cambridge University Press, 550 pp.
Galperin, B., S. Sukoriansky, N. Dikovskaya, P. L. Read, Y. H. Yamazaki, and R. Wordsworth, 2006: Anisotropic turbulence and zonal jets in rotating flows with a β-effect. Nonlinear Processes Geophys., 13, 83–98, https://doi.org/10.5194/npg-13-83-2006.
Held, I. M., 1999: The macroturbulence of the troposphere. Tellus, 51A, 59–70, https://doi.org/10.3402/tellusa.v51i1.12306.
Held, I. M., 2019: 100 years of progress in understanding the general circulation of the atmosphere. A Century of Progress in Atmospheric and Related Sciences: Celebrating the American Meteorological Society Centennial, Meteor. Monogr., No. 59, Amer. Meteor. Soc., https://doi.org/10.1175/AMSMONOGRAPHS-D-18-0017.1.
Held, I. M., and V. D. Larichev, 1996: A scaling theory for horizontally homogeneous, baroclinically unstable flow on a beta plane. J. Atmos. Sci., 53, 946–952, https://doi.org/10.1175/1520-0469(1996)053<0946:ASTFHH>2.0.CO;2.
Jansen, M. F., A. J. Adcroft, R. Hallberg, and I. M. Held, 2015: Parameterization of eddy fluxes based on a mesoscale energy budget. Ocean Modell., 92, 28–41, https://doi.org/10.1016/j.ocemod.2015.05.007.
Kong, H., and M. F. Jansen, 2017: The eddy diffusivity in barotropic β-plane turbulence. Fluids, 2, 54, https://doi.org/10.3390/fluids2040054.
Lapeyre, G., and I. M. Held, 2003: Diffusivity, kinetic energy dissipation, and closure theories for the poleward eddy heat flux. J. Atmos. Sci., 60, 2907–2916, https://doi.org/10.1175/1520-0469(2003)060<2907:DKEDAC>2.0.CO;2.
Larichev, V. D., and I. M. Held, 1995: Eddy amplitudes and fluxes in a homogeneous model of fully developed baroclinic instability. J. Phys. Oceanogr., 25, 2285–2297, https://doi.org/10.1175/1520-0485(1995)025<2285:EAAFIA>2.0.CO;2.
Mak, J., J. R. Maddison, D. P. Marshall, and D. R. Munday, 2018: Implementation of a geometrically informed and energetically constrained mesoscale eddy parameterization in an ocean circulation model. J. Phys. Oceanogr., 48, 2363–2382, https://doi.org/10.1175/JPO-D-18-0017.1.
Mak, J., A. Avdis, T. David, H. S. Lee, Y. Na, Y. Wang, and F. E. Yan, 2022: On constraining the mesoscale eddy energy dissipation time-scale. J. Adv. Model. Earth Syst., 14, e2022MS003223, https://doi.org/10.1029/2022MS003223.
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, https://doi.org/10.1016/j.ocemod.2010.02.001.
Panetta, R. L., 1993: Zonal jets in wide baroclinically unstable regions: Persistence and scale selection. J. Atmos. Sci., 50, 2073–2106, https://doi.org/10.1175/1520-0469(1993)050<2073:ZJIWBU>2.0.CO;2.
Pavan, V., and I. M. Held, 1996: The diffusive approximation for eddy fluxes in baroclinically unstable jets. J. Atmos. Sci., 53, 1262–1272, https://doi.org/10.1175/1520-0469(1996)053<1262:TDAFEF>2.0.CO;2.
Rhines, P. B., 1975: Waves and turbulence on a beta-plane. J. Fluid Mech., 69, 417–443, https://doi.org/10.1017/S0022112075001504.
Rhines, P. B., 1977: The dynamics of unsteady currents. The Sea, Vol. 6, E. D. Goldberg, Ed., Wiley-Interscience, 189–318.
Salmon, R., 1980: Baroclinic instability and geostrophic turbulence. Geophys. Astrophys. Fluid Dyn., 15, 167–211, https://doi.org/10.1080/03091928008241178.
Salmon, R., 1998: Lectures on Geophysical Fluid Dynamics. Oxford University Press, 400 pp.
Scott, R. K., and D. G. Dritschel, 2012: The structure of zonal jets in geostrophic turbulence. J. Fluid Mech., 711, 576–598, https://doi.org/10.1017/jfm.2012.410.
Sukoriansky, S., N. Dikovskaya, and B. Galperin, 2007: On the arrest of inverse energy cascade and the Rhines scale. J. Atmos. Sci., 64, 3312–3327, https://doi.org/10.1175/JAS4013.1.
Thompson, A. F., 2010: Jet formation and evolution in baroclinic turbulence with simple topography. J. Phys. Oceanogr., 40, 257–278, https://doi.org/10.1175/2009JPO4218.1.
Thompson, A. F., and W. R. Young, 2006: Scaling baroclinic eddy fluxes: Vortices and energy balance. J. Phys. Oceanogr., 36, 720–738, https://doi.org/10.1175/JPO2874.1.
Thompson, A. F., and W. R. Young, 2007: Two-layer baroclinic eddy heat fluxes: Zonal flows and energy balance. J. Atmos. Sci., 64, 3214–3231, https://doi.org/10.1175/JAS4000.1.
Vallis, G. K., 2017: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 946 pp.
Vallis, G. K., 2021: Turbulence theory: Imperfect, but necessary. AGU Adv., 2, e2021AV000523, https://doi.org/10.1029/2021AV000523.
Vallis, G. K., and M. E. Maltrud, 1993: Generation of mean flows and jets on a beta plane and over topography. J. Phys. Oceanogr., 23, 1346–1362, https://doi.org/10.1175/1520-0485(1993)023<1346:GOMFAJ>2.0.CO;2.
Wunsch, C., and R. Ferrari, 2018: 100 years of the ocean general circulation. A Century of Progress in Atmospheric and Related Sciences: Celebrating the American Meteorological Society Centennial, Meteor. Monogr., No. 59, Amer. Meteor. Soc., https://doi.org/10.1175/AMSMONOGRAPHS-D-18-0002.1.
Zurita-Gotor, P., 2007: The relation between baroclinic adjustment and turbulent diffusion in the two-layer model. J. Atmos. Sci., 64, 1284–1300, https://doi.org/10.1175/JAS3886.1.