Hostname: page-component-788cddb947-tr9hg Total loading time: 0 Render date: 2024-10-09T12:29:14.135Z Has data issue: false hasContentIssue false

The catalytic effect of near-inertial waves on $\beta$-plane zonal jets

Published online by Cambridge University Press:  04 May 2023

Lin-Fan Zhang
Affiliation:
Department of Mechanics and Engineering Science at College of Engineering, State Key Laboratory for Turbulence and Complex Systems, Beijing 100871, PR China
Jin-Han Xie*
Affiliation:
Department of Mechanics and Engineering Science at College of Engineering, State Key Laboratory for Turbulence and Complex Systems, Beijing 100871, PR China Joint Laboratory of Marine Hydrodynamics and Ocean Engineering, Laoshan Laboratory, Shandong 266237, PR China
*
Email addresses for correspondence: [email protected]

Abstract

Zonal jets and inertia–gravity waves are ubiquitous on planets such as Earth, Jupiter and Saturn. Motivated by the modification of energy flux of balanced flow by inertia–gravity waves, this paper studies the impact of near-inertial waves (NIWs) on zonal jets on a $\beta$-plane. Using a two-dimensional quasi-geostrophic and NIW coupled system on a $\beta$-plane (Xie & Vanneste, J. Fluid Mech., vol. 774, 2015, pp. 143–169), we find NIWs catalytically impact several features of zonal jets. The NIWs inhibit jet formation due to the waves’ catalytic induction of downscale mean energy flux. As the strength of NIWs increases, a critical point exists beyond which zonal jets are annihilated. The jet spacing is captured by the Rhines scale $L\sim \sqrt {U/\beta }$ with $U$ estimated from the upscale energy flux induced by the mean flow alone, which again shows that the NIWs’ impact is catalytic. Also, the temporal asymmetry of NIWs leads to the spatial asymmetry of jet dynamics. The jet profiles are asymmetric with a stronger shear on the left flank. And similar to the left turning of vortex dipole under the impact of NIWs, the NIW-modified jets migrate poleward. The NIWs also show a catalytic role in jet migration: the net momentum flux directly induced by NIWs is of secondary importance in the zonal mean momentum dynamics and impedes jet migration, while the advective effect of NIW-modified mean flow dominates the jet migration velocity.

Type
JFM Papers
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Rotation and stratification strongly impact geostrophic flows and lead to two distinctive motions: slow large-scale balanced flows and fast small-scale inertia–gravity waves (IGWs) (Smith & Waleffe Reference Smith and Waleffe2002; Vallis Reference Vallis2006). The variation of the Coriolis effect with latitude, the $\beta$-effect, leads to east–west flows with alternating directions, so-called zonal jets, which are ubiquitous on planets such as Jupiter, Saturn and Earth (Feldstein Reference Feldstein1998; Sanchez-Lavega & Rojas Reference Sanchez-Lavega and Rojas2000; García-Melendo Reference García-Melendo2001; Uppala et al. Reference Uppala2005). On Jupiter, the zonation of clouds shows the existence of jet streams (Vasavada & Showman Reference Vasavada and Showman2005). On Earth, zonal jets play an important role in transporting energy, momentum and scalar quantities such as heat and salt, which impact the pattern of atmospheric and oceanic flows and living creatures (Rhines Reference Rhines1994). Nonlinear interactions between IGWs and mean flows have long been studied in geostrophic fluid dynamics (Lighthill Reference Lighthill1978; Bühler Reference Bühler2009). Following this spirit, this paper focuses on the impact of IGWs on the formation, spacing, profile and dynamics of $\beta$-plane zonal jets.

Jet formation can be understood from turbulence phenomenology by matching the characteristic time scales of inverse cascade (Kraichnan Reference Kraichnan1967) and Rossby waves, whose anisotropic dispersion relation leads to an anisotropic barrier of upscale energy flux and favours jets (Vallis & Maltrud Reference Vallis and Maltrud1993). In a weakly nonlinear regime, Manfroi & Young (Reference Manfroi and Young1999) showed that large-scale instability leads to the formation of slowly evolving jets. In a baroclinic model, Berloff, Kamenkovich & Pedlosky (Reference Berloff, Kamenkovich and Pedlosky2009) found that jets form through the secondary instability of the baroclinically unstable primary modes. From the points of the material conservation of potential vorticity, zonal jets arise inevitably because of the quasi-horizontal mixing of the background gradient of potential vorticity by nonlinear eddy motions over limited latitudinal regions (Dritschel & McIntyre Reference Dritschel and McIntyre2008; Scott & Dritschel Reference Scott and Dritschel2012; Galperin & Read Reference Galperin and Read2019). In the quasi-linear approximation, where the inverse cascade is absent, the mechanisms of stochastic structural instability (Farrell & Ioannou Reference Farrell and Ioannou2003) and zonostrophic instability (Srinivasan & Young Reference Srinivasan and Young2012) are proposed to explain jet formation.

What determines jet spacing is a long-lasting question. The Rhines scale (Rhines Reference Rhines1975) $L_{Rh}=\sqrt {U/\beta }$, where $U$ is the root-mean-square (r.m.s.) velocity, is found to determine jet spacing in many circumstances (Williams Reference Williams1978; Rhines Reference Rhines1979, Reference Rhines1994; Danilov & Gurarie Reference Danilov and Gurarie2000; Galperin et al. Reference Galperin, Sukoriansky, Dikovskaya, Read, Yamazaki and Wordsworth2006; Chemke & Kaspi Reference Chemke and Kaspi2015a). Nevertheless, there is freedom in choosing the characteristic velocity $U$. For instance, Dritschel & McIntyre (Reference Dritschel and McIntyre2008) used the vortex-peak velocity and Scott & Dritschel (Reference Scott and Dritschel2012) calculated the r.m.s. velocity from the energy balance between the upscale energy flux of two-dimensional (2-D) turbulence and the frictional dissipation. Based on the characteristic scales of inverse cascade and Rossby wave, Maltrud & Vallis (Reference Maltrud and Vallis1991) introduced the spectral Rhines scale, which describes the transition scale between turbulent and wavelike regimes. Dunkerton & Scott (Reference Dunkerton and Scott2008) calculated the geometrical Rhines scale from the conservation of absolute angular momentum on a sphere for a jet's latitudinal spacing. However, the Rhines scale does not always predict the jet spacing when other characteristic scales such as the Rossby deformation radius $L_{D}$ exist (Okuno & Masuda Reference Okuno and Masuda2003; Smith Reference Smith2004; Dritschel & McIntyre Reference Dritschel and McIntyre2008).

Zonal jets have interesting asymmetric velocity profiles. García-Melendo (Reference García-Melendo2001) and Sánchez-Lavega et al. (Reference Sánchez-Lavega2008) observed asymmetric zonal jets on Jupiter, where the eastward jets form cusps at the maximum velocity while westward flows are smooth. This asymmetry is explained by the diagnostic relation between zonal velocity and potential vorticity staircases (Marcus & Lee Reference Marcus and Lee1998; Dritschel & McIntyre Reference Dritschel and McIntyre2008). Woillez & Bouchet (Reference Woillez and Bouchet2019) asymptotically derived the eastward zonal jets’ velocity profile on a $\beta$-plane, and they showed that the potential vorticity staircase is an idealized approximation and the actual jet velocity profile is also controlled by the effects of friction, forcing and $\beta$. Lemasquerier, Favier & Le Bars (Reference Lemasquerier, Favier and Le Bars2021) experimentally found that zonal jets with a weak forcing can exist instantaneously even without the process of potential vorticity mixing.

Among the rich dynamics of zonal jets (cf. Galperin & Read Reference Galperin and Read2019; Lemasquerier et al. Reference Lemasquerier, Favier and Le Bars2021), one interesting phenomenon is their meridional migration. In the atmosphere, jets are found to migrate poleward in both observations (Riehl, Yeh & La Seur Reference Riehl, Yeh and La Seur1950; Dickey, Marcus & Hide Reference Dickey, Marcus and Hide1992) and numerical simulations (James & Dodd Reference James and Dodd1996; Chemke & Kaspi Reference Chemke and Kaspi2015b), while jets’ equatorward migration was observed in ocean models (Chan, Plumb & Cerovecki Reference Chan, Plumb and Cerovecki2007; Ashkenazy & Tziperman Reference Ashkenazy and Tziperman2016). Using a semi-hemisphere zonally re-entrant ocean model, Chan et al. (Reference Chan, Plumb and Cerovecki2007) found the primary jet oscillates meridionally and inconspicuously propagates poleward over a short range, and the secondary jets systematically migrate equatorward slowly from high latitudes and eventually merge with the primary jet. Besides, Williams (Reference Williams2003) described a similar equatorward jet migration in a model of Jupiter's atmosphere. Apart from gradual migration, abrupt jet transitions are also observed in forced-dissipative 2-D quasi-geostrophic (QG) systems (Bouchet & Venaille Reference Bouchet and Venaille2012; Rolland & Simonnet Reference Rolland and Simonnet2015; Bouchet, Marston & Tangarife Reference Bouchet, Marston and Tangarife2018; Bouchet, Rolland & Simonnet Reference Bouchet, Rolland and Simonnet2019).

Among the wide range of frequency of IGWs, near-inertial waves (NIWs) contain around $50\,\%$ of the total wave energy in the ocean (Fu Reference Fu1981; Ferrari & Wunsch Reference Ferrari and Wunsch2009); therefore, we pay special attention to the effects of NIWs on zonal jets in this paper. The interaction between NIWs and balanced flow is twofold. When propagating in balanced flows, NIWs are affected by advection, refraction and dispersion, and they are trapped by anticyclones of balanced flow (Kunze, Schmitt & Toole Reference Kunze, Schmitt and Toole1995; Elipot, Lumpkin & Prieto Reference Elipot, Lumpkin and Prieto2010; Joyce et al. Reference Joyce, Toole, Klein and Thomas2013; Thomas et al. Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020). Because the negative vorticity shifts the lower bound of the internal waveband to a frequency slightly below the Coriolis frequency $f$, NIWs with frequency less than $f$ concentrate in anticyclones (Kunze Reference Kunze1985). The NIWs trapped in anticyclones are also found to propagate downward to the deeper ocean, which is called the inertial chimney phenomenon (Lee & Niiler Reference Lee and Niiler1998). Considering the comparable horizontal scales of NIWs and balanced flow, Young & Ben Jelloul (Reference Young and Ben Jelloul1997) asymptotically derived an amplitude equation (YBJ equation) of first-order accuracy in the Burger number of the NIWs. Asselin & Young (Reference Asselin and Young2019) further proposed the $\text {YBJ}^+$ equation, which captures the second-order effect of the linear dispersion relation of IGWs. Danioux, Vanneste & Bühler (Reference Danioux, Vanneste and Bühler2015) applied the YBJ model to explain the concentration of NIWs in anticyclones based on conserved quantities. Using the three-dimensional YBJ equation, Asselin et al. (Reference Asselin, Thomas, Young and Rainville2020) studied the weakly dispersive NIWs’ wavevector impacted by refraction and strain, which is justified by observation (Thomas et al. Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020).

With large horizontal and small vertical scales, NIWs are weakly dispersive and potentially have a strong impact on the balanced flow. With $O(1)$ Rossby number, Gertz & Straub (Reference Gertz and Straub2009) and Taylor & Straub (Reference Taylor and Straub2016) numerically found that NIWs provide an effective energy sink for wind-driven ocean gyres. Spontaneous generation of IGWs (cf. Vanneste (Reference Vanneste2013), and references therein) is thought to be important for energy sink for the balanced flow, which is exponentially weak in the regime of small Rossby number (Vanneste Reference Vanneste2008), and it can be effective when the Rossby number is of $O(1)$, which was shown in numerical simulations (Danioux et al. Reference Danioux, Vanneste, Klein and Sasaki2012; Plougonven & Snyder Reference Plougonven and Snyder2007) and by analytical solutions (Zeitlin Reference Zeitlin2008; Thomas Reference Thomas2012; Whitt & Thomas Reference Whitt and Thomas2015; Grisouard & Thomas Reference Grisouard and Thomas2016). However, the strength of the NIW sink may be overestimated, as a considerable amount of wave energy can be reabsorbed by the mean flow (Nagai et al. Reference Nagai, Tandon, Kunze and Mahadevan2015; Shakespeare & McC. Hogg Reference Shakespeare and McC. Hogg2018). Stimulated loss of balance (Xie & Vanneste Reference Xie and Vanneste2015; Wagner & Young Reference Wagner and Young2016; Rocha, Wagner & Young Reference Rocha, Wagner and Young2018) is also proposed as a mechanism to absorb energy from balanced flow by NIWs. Based on a generalized Lagrangian mean framework (Andrews & Mcintyre Reference Andrews and Mcintyre1978; Soward & Roberts Reference Soward and Roberts2010; Salmon Reference Salmon2013), this mechanism implies that the decrease of NIW horizontal scale by mean-flow advection and refraction leads to energy conversion from balanced motions to NIWs. Weakened balanced flows, in turn, enfeeble wave refraction and thus halt the drainage of waves into the interior ocean (Asselin & Young Reference Asselin and Young2020). Using the coupled QG–YBJ model (Xie & Vanneste Reference Xie and Vanneste2015) with initially uniform NIWs, Kafiabad, Vanneste & Young (Reference Kafiabad, Vanneste and Young2021) calculated the NIW frequency shift that is linearly proportional to wave kinetic energy.

The presence of NIWs also catalyses other processes in mean flows. With linear analytical solutions, Thomas (Reference Thomas2012) found that IGWs play a catalytic role when fronts lose balance to ageostrophic motions in frontogenesis. By analysing frequency-space filtered numerical data, Barkan, Winters & McWilliams (Reference Barkan, Winters and McWilliams2017) and Barkan et al. (Reference Barkan, Srinivasan, Yang, McWilliams, Gula and Vic2021) found that NIWs catalyse energy transfer from mesoscale to submesoscale. By studying a Galerkin-truncated 2-D model of rotating stratified flow, Thomas & Arun (Reference Thomas and Arun2020) found that high-energy NIWs can transfer energy to barotropic flow and impel a forward energy cascade of the balanced flows that assists the forward energy cascade of waves in turn. By studying a 2-D model derived asymptotically in Xie & Vanneste (Reference Xie and Vanneste2015), in statistically steady states with external forcing and dissipation, Xie (Reference Xie2020) proposed that based on the conservations of energy, wave action and potential enstrophy, NIWs catalyse a downscale energy transfer of QG mean flows.

This paper studies the impact of NIWs on jet dynamics on a $\beta$-plane, which is motivated by ubiquitous waves and jets and the impact of NIWs on the energy flux of mean flows. The rest of this paper is organized as follows. In § 2, we first present the coupled model of QG mean flows and NIWs following Xie & Vanneste (Reference Xie and Vanneste2015). In § 3, we show the impact of NIWs on the energy evolution, the energy spectra of mean flow and NIWs and qualitatively analyse the catalytic effect of NIWs from the perspective of the mean-flow energy flux. In § 4, we quantitatively study the impact of NIWs on jet formation and find that NIWs weaken jets. In § 5, we study the dependence of jet spacing on NIW strength. A modified Rhines scale estimated using the advective upscale mean energy flux alone captures the jet spacing well. This indirect impact of NIWs again shows a catalytic effect. In § 6, we find that NIWs break jets’ spatial symmetry by introducing asymmetric jet flank and leading to jets’ poleward migration. In studying the jet migration from the mean momentum balance, the NIWs’ direct impact is much weaker than that of the asymmetric mean flow, which shows the third NIW catalytic effect. Finally, in § 7, we summarize and discuss our results.

2. Coupled model of NIW–mean flow interaction on a $\beta$-plane

For simplicity and numerical efficiency and to capture the key feature of wave–mean flow interaction, we study the interaction between a single-vertical-mode NIW and a barotropic QG mean flow using the 2-D model derived by Xie & Vanneste (Reference Xie and Vanneste2015) on a $\beta$-plane:

(2.1a)$$\begin{gather} \partial_t q+{J}(\psi,q)=0 , \end{gather}$$
(2.1b)$$\begin{gather}\partial_t\phi+{J}(\psi,\phi)+\mathrm{i} g(y)\phi-\frac{\mathrm{i} f_0}{2}{L}^2\phi+\frac{\mathrm{i}}{2}\nabla^2\psi\phi=0, \end{gather}$$

where

(2.2)\begin{equation} q=\nabla^2\psi+ g(y) +\frac{\mathrm{i} f_0}{2}{J}(\phi^*,\phi)+\frac{f_0}{4}\nabla^2|\phi|^2\end{equation}

is the NIW-modified quasi-geostrophic potential vorticity (QGPV). Here, $\psi$ is the Lagrangian mean streamfunction and the horizontal velocity of QG mean flows is $(u,v)=(-\partial _y \psi, \partial _x \psi )$; $f_0=\text {const.}$ is the local Coriolis frequency; and $\phi$ is the complex wave amplitude and links to the two horizontal velocities $(u_0,v_0)$ through $u_0+\mathrm {i} v_0=-\mathrm {i} f_0 \phi \, \mathrm {e}^{-\mathrm {i} f_0 t+\mathrm {i} m z}$ with $m$ the vertical wavenumber of the single-vertical-mode NIW. The asterisk denotes the complex conjugate and ${J}(a,b)=a_xb_y-a_yb_x$ is the Jacobian. Also, $g(y)$ is the variation of the Coriolis parameter $f$ in the meridional direction such that $f=f_0+g(y)$; therefore, the $\beta$-effect is captured by $\beta =\partial _y g$. This model consists of the equation describing the material-invariant QGPV (2.1a) and the linear equation capturing the slow modulation of NIW amplitude (Young & Ben Jelloul Reference Young and Ben Jelloul1997). In (2.1b) we consider a modified dispersion (Asselin & Young Reference Asselin and Young2019) with the operator

(2.3)\begin{equation} {L}^2=\frac{\nabla^2}{\dfrac{f_0^2}{N^2}m^2-\dfrac{1}{4}\nabla^2}, \end{equation}

where $N$ is the buoyancy frequency. Here, the Laplacian operator in the denominator should be understood in the spectral space. This modified dispersion term captures the high-order corrections of the NIW dispersion relation and improves the numerical efficiency by bounding the frequency as the wavenumber increases.

For simplicity, we introduce a modified meridional-dependent $\beta$-effect, $\beta (y)$, which is periodic in $y$ instead of the traditional choice of a constant, to perform numerical simulations in a periodic domain. Practically, to extend the region with a constant $\beta$, we prescribe

(2.4) \begin{equation} \beta(y)= \begin{cases} \beta_0 \dfrac{\mathrm{e}^{l_b y}-\mathrm{e}^{{-}l_b y}}{\mathrm{e}^{l_b y}+\mathrm{e}^{{-}l_b y}}, & 0\leq y < L_y/4 ,\\ \beta_0 \dfrac{\mathrm{e}^{l_b (L_y/2-y)}-\mathrm{e}^{{-}l_b (L_y/2-y)}}{\mathrm{e}^{l_b (L_y/2-y)}+\mathrm{e}^{{-}l_b (L_y/2-y)}}, & L_y/4\leq y < 3L_y/4 ,\\ \beta_0 \dfrac{\mathrm{e}^{l_b ({-}L_y+y)}-\mathrm{e}^{{-}l_b ({-}L_y+y)}}{\mathrm{e}^{l_b ({-}L_y+y)}+\mathrm{e}^{{-}l_b ({-}L_y+y)}}, & 3L_y/4\leq y < L_y , \end{cases} \end{equation}

where $\beta _0=\text {const.}$, $l_b$ is a constant that controls the range of the variation in $\beta$ and $L_y$ is the domain size in the meridional direction. Our analysis uses the data obtained from the region $\approx [1/l_b, L_y/2-1/l_b]$, where $\beta$ is almost a constant. The other half of the domain is only introduced for numerical simulation, and it is unphysical because $\beta$ should be positive in the northern hemisphere. We use the current periodic set-up mainly because of its simplicity and computational efficiency. The expressions of $\beta (y)$ and $g(y)$ are illustrated in figure 1. An alternative set-up is to consider a constant-$\beta$ system in a channel, which is widely used. For example, Abernathey, Marshall & Ferreira (Reference Abernathey, Marshall and Ferreira2011) numerically simulated a $\beta$-plane channel with a sponge layer at the northern boundary to study the Antarctic Circumpolar Current.

Figure 1. Illustration of (a) $\beta (y)$ and (b) $g(y)$ with $L_y=2{\rm \pi}$ and $l_b=10$ in (2.4).

The coupled system (2.1) conserves energy, potential enstrophy and wave action, which reads

(2.5a) $$\begin{gather} \mathcal{E}=\int \left\{\underset{\mathcal{E}_{QG}}{\underbrace{\frac{1}{2}|\boldsymbol{\nabla} \psi|^2}}+ \underset{\mathcal{E}_{\beta}}{\underbrace{\frac{1}{2}f_0g(y) |\phi|^2}}+ \underset{\mathcal{E}_{NIW}}{\underbrace{\frac{f_0^2}{4} |L\phi|^2}} \right\}\mathrm{d}\boldsymbol{x}, \end{gather}$$
(2.5b)$$\begin{gather}\mathcal{P}=\int q^2\,\mathrm{d}\boldsymbol{x}, \end{gather}$$
(2.5c)$$\begin{gather}\mathcal{A}=\int |\phi|^2\,\mathrm{d}\boldsymbol{x}. \end{gather}$$

Here we decompose the total energy into three components corresponding to the mean, $\beta$-effect and wave potential energy for convenience of later discussion.

In this paper, we focus on the statistically steady states of the NIW–mean flow coupled system with external forcing and dissipation. Following Xie (Reference Xie2020), for a straightforward quantification of energy injection rate to the mean and wave components, forcing and dissipation are added to the mean vorticity equation instead of the QGPV equation and the wave amplitude equation. Therefore we obtain the system

(2.6a)$$\begin{gather} \partial_t\nabla^2\psi+\mathcal{N}(\psi,\phi) = {D}_1\nabla^2\psi+M_{F}k_{f}^{1/2}\mathsf{F}_1, \end{gather}$$
(2.6b)$$\begin{gather}\partial_t\phi+{J}(\psi,\phi)+\mathrm{i} g(y)\phi-\frac{\mathrm{i} f_0}{2}L^2\phi+\frac{\mathrm{i}}{2}\nabla^2\psi\phi = {D}_2\phi+RM_{F}\frac{m}{2N k_{f}^{3/2}}\mathsf{F}_2, \end{gather}$$

where the artificial dissipation operators take the form of ${D}_1=-\alpha _1+\nu _1\nabla ^6$ and ${D}_2=\alpha _2\nabla ^{-2}+\nu _2\nabla ^6$ with constants $\alpha _i$ and $\nu _i$ ($i=1,2$). The external forcings $\mathsf {F}_i$ ($i=1,2$), which are white-noise in time, are isotropic in space and centre around wavenumber $|\boldsymbol {k}|=k_{f}$, where $\boldsymbol {k}=(k,l)$ is the 2-D wavenumber vector in spectral space. Therefore,

(2.7)\begin{equation} \langle\mathsf{F}_i(\boldsymbol{x}_1,t_1) \mathsf{F}_i(\boldsymbol{x}_2,t_2)\rangle= C_0{J}_0(k_{f}|\boldsymbol{x_1}-\boldsymbol{x_2}|)\delta(t_1-t_2), \end{equation}

where $\langle \cdot \rangle$ is an ensemble average, $C_0$ is the covariance and ${J}_0$ is the zeroth-order Bessel function. Here $M_{F}$ and $R$ control the external forcing magnitude and $k_{f}^{1/2}$ and $m/(Nk^{3/2}_{f})$ are the normalized coefficients. Combining (2.5a) and (2.6), we estimate the energy injection rates of QG mean flow and NIW as $\epsilon _{QG}\sim M_{F}^2/2$ and $\epsilon _{NIW}\sim R^2M_{F}^2f_0^2m^2/(16f_0^2m^2+4N^2k_{f}^2)$. Therefore, we define the ratio between energy injection rates of NIW and mean flow as

(2.8)\begin{equation} \lambda =\frac{R^2m^2f_0^2}{8m^2f_0^2+2N^2k_{f}^2}.\end{equation}

Here, the denominator is consistent with the modification by Asselin & Young (Reference Asselin and Young2019), and when this modification is absent, $\lambda = R^2/8$.

We decompose the nonlinear term $\mathcal {N}(\psi,\phi )$ in (2.6a) into different components for convenience in later discussions as

(2.9)\begin{equation} \mathcal{N}(\psi,\phi)=\mathcal{N}_{\beta}(\psi)+ \mathcal{N}_m(\psi)+\mathcal{N}_{W}(\psi,\phi),\end{equation}

where

(2.10a)$$\begin{gather} \mathcal{N}_{\beta}(\psi)=g_y\psi_x+\frac{f_0}{2}g_y(\phi^*_x\phi +\phi^*\phi_x), \end{gather}$$
(2.10b)$$\begin{gather}\mathcal{N}_m(\psi)={J}(\psi,\nabla^2\psi), \end{gather}$$
(2.10c)$$\begin{gather}\mathcal{N}_{W}(\psi,\phi)=\frac{- f_0^2}{4}({J}(\phi^*,L^2\phi) - {J}(L^2\phi^*,\phi)) -\frac{\mathrm{i} f_0^2}{8}\nabla^2(\phi L^2\phi^*-\phi^*L^2\phi)\nonumber\\ -\frac{f_0}{2}\boldsymbol{\nabla} \boldsymbol{\cdot} {J}(\boldsymbol{\nabla} \psi,|\phi|^2). \end{gather}$$

3. Basic properties of the NIW–QG coupled system on a $\beta$-plane

We run the numerical simulations using a Fourier pseudospectral method with $2/3$ dealiasing in space and a resolution of $512\times 512$ in a domain size of $2{\rm \pi} \times 2{\rm \pi}$. We apply a fourth-order explicit Runge–Kutta temporal scheme (cf. Cox & Matthews Reference Cox and Matthews2002) where the linear terms are solved by an integrating factor method, and the nonlinear terms are explicitly approximated. We choose $\beta _0=80$ and $f_0 m/N=32$ in all simulations. The QG mean flows and NIWs are driven by external forcing at a forcing wavenumber $k_{f}=32$. We vary two parameters $M_{F}$ and $R$ to create a matrix of results. Thus, the energy injection rates are $\epsilon _{QG}=8\times 10^{-5},3.2\times 10^{-4},7.2\times 10^{-4}$ with $M_{F}=1000,2000,3000$. Parameter $R$ ranges from $0$ to $3.6$, and the energy injection ratio of NIWs to mean flows $\lambda$ varies from $0$ to $1.296$ by (2.8).

We first run numerical simulations without NIWs as references to study the wave effect. Figure 2 shows a snapshot of the vorticity field $\nabla ^2\psi$ of a NIW-free simulation with $\epsilon _{QG}=8\times 10^{-5}$ at a statistically steady state when jets form. Then we use this vorticity field as the initial condition for other NIW-involved simulations with non-zero $\lambda$.

Figure 2. A snapshot of QG mean flow vorticity fields in the statistically steady state of the NIW-free simulation with $\epsilon _{QG}=8\times 10^{-5}$.

In this section, we focus on qualitative descriptions of characteristic features and leave quantitative studies to §§ 46.

3.1. Energy evolution and statistically steady states

In this subsection, we run the coupled system into statistically steady states. We display NIWs’ impact on the total, wave and mean energy evolution. Figure 3(a) shows the evolution of mean energy with different energy injection ratios $\lambda$. When NIWs are injected, the mean energy $\mathcal {E}_{QG}$ decreases and gradually reaches a statistically steady state. With a stronger NIW energy injection rate, the steady value of $\mathcal {E}_{QG}$ becomes weaker, and the relaxation time to reach the statistically steady states increases and then decreases. This non-monotonic dependence implies the existence of a phase transition associated with the disappearance of the jet (cf. figure 4d). In figure 3(b), the $\beta$-related NIW energy, $\mathcal {E}_{\beta }$, oscillates around zero and the amplitude of oscillation increases as $\lambda$ increases. Compared with $\mathcal {E}_{QG}$, $\mathcal {E}_{\beta }$ is relatively small. Figure 3(c) shows that the evolution of wave potential energy $\mathcal {E}_{NIW}$ has three stages. First, there is a sudden wave energy adjustment at a short time interval, then the wave energy increases in a comparable time scale to the mean energy changes and, finally, a statistically steady state is reached. Since $\mathcal {E}_{\beta }$ and $\mathcal {E}_{NIW}$ are small compared with $\mathcal {E}_{QG}$, figure 3(d) shows that the total energy $\mathcal {E}$ is dominated by $\mathcal {E}_{QG}$ but with stronger fluctuations.

Figure 3. Evolutions of (a) QG mean, (b) $\beta$-related, (c) $\mathrm {NIW}$ and (d) total energy (cf. (2.5a)) in simulations with $\lambda =0.016, 0.144, 0.4, 0.784, 1.296$. In (b), for clarity we vertically shift the curves following $\mathcal {E}_{\beta }(\lambda =0.144)+0.02$, $\mathcal {E}_{\beta }(\lambda =0.4)+0.05$, $\mathcal {E}_{\beta }(\lambda =0.784)+0.1$, $\mathcal {E}_{\beta }(\lambda = 1.296)+0.15$. The legend in (a) applies to all four panels.

Figure 4. Snapshots of QG mean vorticity fields, $\nabla ^2\psi$, at statistically steady states with (a) $\lambda =0.144$, (b) $\lambda =0.4$, (c) $\lambda =0.784$ and (d) $\lambda =1.296$. The mean energy injection rate is $\epsilon _{QG}=8\times 10^{-5}$.

Snapshots of statistically steady QG mean flow vorticity field $\nabla ^2\psi$ are presented in figure 4. When the energy injection rate of NIWs is weak (figure 4a with $\lambda =0.144$), zonal jets remain similar in magnitude and shape compared with those in the NIW-free case. As the energy injection ratio of NIW to mean flows increases ($\lambda =0.4$ and $0.784$ shown in figures 4b and 4c), jets weaken and the jet spacing increases. When the energy injection ratio is large enough (figure 4d with $\lambda =1.296$), jets disappear. Compared with the vorticity field without NIWs in figure 2, zonal jets are weakened by NIWs and their geometries are modified, which we study quantitatively in §§ 46.

3.2. Energy spectra

To present the impact of NIWs on the spectrum of QG mean flow and, in particular, the jets, based on the 2-D energy spectrum

(3.1)\begin{equation} E(\boldsymbol{k}) =\frac{1}{2}|\boldsymbol{k}|^2 \hat{\psi}(\boldsymbol{k}) \hat{\psi}^*(\boldsymbol{k}), \end{equation}

where the hat $\hat {\cdot }$ denotes the Fourier transform, we define the one-dimensional spectra of the total mean energy, the jet energy and the perturbation energy as

(3.2a)$$\begin{gather} E_T(K) = \sum_{|\boldsymbol{k}|=K-1/2}^{K+1/2} E(\boldsymbol{k}), \end{gather}$$
(3.2b)$$\begin{gather}E_Z(K) = \sum_{|\boldsymbol{k}|=K-1/2}^{K+1/2} E(0,l), \end{gather}$$
(3.2c)$$\begin{gather}E_R(K) = \sum_{|\boldsymbol{k}|=K-1/2}^{K+1/2} E(k\neq0,l). \end{gather}$$

Here, $E_T(K)=E_Z(K)+E_R(K)$. In the classic 2-D $\beta$-plane turbulence, the jet and perturbation energy spectra follow the scalings $E_Z(K)=C_Z\beta ^2K^{-5}$ and $E_R(K)=C_K\epsilon ^{2/3}K^{-5/3}$, where $\epsilon$ is the viscous dissipation rate (Galperin, Sukoriansky & Dikovskaya Reference Galperin, Sukoriansky and Dikovskaya2010). The 2-D potential energy spectrum of NIWs is defined as

(3.3)\begin{equation} E_W(K) =\frac{f_0^2}{4}\sum_{|\boldsymbol{k}|=K-1/2}^{K+1/2} \widehat{{L}\phi}(\boldsymbol{k}) \widehat{{L}\phi}^*(\boldsymbol{k}). \end{equation}

First, we show the spectra of the statistically steady state in the NIW-free simulation in figure 5. At scales between the spectral peak and the forcing wavenumber, the mean and perturbation energy spectra follow $K^{-5/3}$ and $K^{-4}$ scalings, respectively. This $K^{-4}$ scaling was derived by Danilov & Gurarie (Reference Danilov and Gurarie2004) with consideration of a sawtooth profile of zonal mean vorticity, making it different from the classic $K^{-5}$ scaling proposed by Rhines (Reference Rhines1975). To show the wave effect, we present energy spectra with different energy injection ratios $\lambda$ in figure 6. From figure 6(a), we find that as $\lambda$ increases from $0$ to $1.024$, the peaks of $E_Z(K)$ move to smaller wavenumber; in the intermediate scale between spectral peak and the energy injection scale, the jet energy drops but keeps the $K^{-4}$ scaling. However, when $\lambda =1.296$, jets disappear and $E_Z(K)$ deviates from the classic $K^{-5}$ scaling. Figure 6(b) shows that as the NIW energy injection rate gradually becomes comparable with the mean energy injection rate, the perturbation energy drops at the large scale and increases at the small scale; the peak of $E_R(K)$ also moves to larger wavenumbers; and the range of the perturbation energy spectrum with the $K^{-5/3}$ scaling shortens and finally disappears. When jets completely disappear with $\lambda = 1.296$, for wavenumbers smaller than $k_{f}$, $E_R(K)$ follows a $K^{}$ scaling. The combination of jet and perturbation energy spectra, $E_Z(K)$ and $E_R(K)$, respectively, explains the variation of total energy spectrum scaling from $K^{-4}$ to $K^{-5/3}$ and finally to $K^{}$ under the influence of increasing energy injection ratio $\lambda$, which is shown in figure 6(c). However, when the NIW energy injection rate is large enough, such as $\lambda = 1.296$, figures 6(a)–6(c) show that at scales between the spectral peak and the forcing wavenumber, the jet and perturbation spectra differ from the scalings in the theories without wave impact. As to the potential energy spectrum of NIWs in figure 6(d), as the energy injection ratio $\lambda$ increases, for wavenumbers smaller than the forcing wavenumber, it reaches its largest value when $\lambda =1.024$, corresponding to jet disappearance. If we treat the jet strength as an order parameter and the NIW as a perturbation, the increase in perturbation strength implies the existence of a phase transition, which we study in § 4.

Figure 5. The QG mean energy spectrum without NIWs at a statistically steady state with $\epsilon _{QG}=8\times 10^{-5}$.

Figure 6. The NIW-modified QG mean and NIW energy spectra at statistically steady states with $\lambda$ ranging from $0$ to $1.296$: (a) $E_Z(K)$, (b) $E_R(K)$, (c) $E_T(K)$ and (d) $E_W(K)$. The legend in (c) applies to all four panels. The black dash-dotted lines are power functions for reference. The black arrows show the direction of increasing $\lambda$.

3.3. Energy flux across scales

Since the energy flux of the QG mean flows across scales relates to jet formation (Rhines Reference Rhines1975; Vallis & Maltrud Reference Vallis and Maltrud1993; Chekhlov et al. Reference Chekhlov, Orszag, Sukoriansky, Galperin and Staroselsky1996; Sukoriansky, Dikovskaya & Galperin Reference Sukoriansky, Dikovskaya and Galperin2007), this subsection focuses on the QG mean energy flux influenced by NIWs to explain jet weakening and annihilation observed in vorticity fields and mean energy spectra. The spectral energy flux of QG mean flow is defined as

(3.4)\begin{equation} F_{QG}(K)=\sum_{|\boldsymbol{k}|=0}^{K} \frac{-|\boldsymbol{k}|^2}{2}(\hat{\psi}\hat{\mathcal{N}}^*+{\rm c.c.}), \end{equation}

and the positive and negative values of $F_{QG}$ correspond to downscale and upscale fluxes, respectively.

To show the influence of mean flow and NIWs separately, we decompose the energy flux of mean flow as follows:

(3.5a)$$\begin{gather} F_{QG} = F_{QG}^{m} + F_{QG}^{NIW}, \quad \mathrm{with} \end{gather}$$
(3.5b)$$\begin{gather}F_{QG}^{m}=\sum_{|\boldsymbol{k}|=0}^{K}\frac{1}{2}(\hat{\psi} \hat{\mathcal{N}}^*_{m}+ {\rm c.c.}), \end{gather}$$
(3.5c)$$\begin{gather}F_{QG}^{NIW}=\sum_{|\boldsymbol{k}|=0}^{K}\frac{1}{2}(\hat{\psi} \hat{\mathcal{N}}^*_{NIW}+ {\rm c.c.}), \end{gather}$$

where the superscripts ‘$m$’ and ‘${NIW}$’ denote the effects of mean flow and NIWs, respectively.

When NIWs are absent, the coupled system (2.1) is identical to the 2-D turbulence system on a $\beta$-plane; therefore, the mean energy should transfer upscale with a constant value, which is shown in figure 7. Here, the finite hyperviscosity brings about a weak downscale flux.

Figure 7. The QG mean energy flux without NIWs at a statistically steady state with $k_{f}=32$ and $\epsilon _{QG}=8\times 10^{-5}$. The negative and positive values denote the upscale and downscale energy fluxes, respectively.

In figure 8 we present the energy flux of mean flow with different energy injection rates of NIWs. The NIWs induce bidirectional flux of mean-flow energy, and as $\lambda$ increases, the portion of downscale flux increases, as shown in figures 8(a)–8(c). Therein, both the energy flux induced by mean flow, $F_{QG}^{m}$, and the energy flux induced by NIWs, $F_{QG}^{NIW}$, increase and the cancellation of these two parts leaves a bidirectional mean-flow energy flux. When $\lambda$ is large enough (cf. figure 8d), all the mean energy transfers downscale. The presence of mean-flow inertial ranges with constant energy fluxes above and below the forcing wavenumber implies negligible conversion between QG mean flow and NIWs, which is consistent with the catalytic effect of NIWs on the mean energy flux (Xie Reference Xie2020; Barkan et al. Reference Barkan, Srinivasan, Yang, McWilliams, Gula and Vic2021).

Figure 8. Energy flux of QG mean flow in NIW-containing cases of (a) $\lambda =0.144$, (b) $\lambda =0.4$, (c) $\lambda =0.784$ and (c) $\lambda =1.296$ with $k_{f}=32$ and $\epsilon _{QG}=8\times 10^{-5}$.

4. Dependence of jet magnitude on the strength of NIWs

In this section, we study the dependence of jet strength on the strength of NIWs, which $\lambda$ parametrizes. We define jet energy as $\mathcal {E}_{jet}=\int \bar {u}^2/2\,\mathrm {d} \boldsymbol {x}$, where $\bar {u}$ is the zonal-averaged zonal QG mean velocity. Figure 9(a) shows the dependence of $\mathcal {E}_{jet}/\mathcal {E}_{QG}$ on $\lambda$ at statistically steady states with different mean energy injection rates. When the mean energy injection rate, $\epsilon _{QG}$, is fixed, there exists a critical value $\lambda _c$: when $0<\lambda <\lambda _c$, $\mathcal {E}_{jet}/\mathcal {E}_{QG}$ is almost a constant, and when $\lambda$ goes across the critical points $\lambda _c$, it abruptly drops to a value close to $0$. Therefore, as the NIW energy injection rate increases from $0$, the QG–NIW coupled system experiences a first-order phase transition, across which jets suddenly disappear.

Figure 9. Ratio between jet energy and the QG mean-flow energy with different mean energy injection rates. In (b), the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

Next, we explore the dependence of QG mean energy injection rate $\epsilon _{QG}$ on critical values $\lambda _c$. Since near the critical point the jet magnitude is small, in the mean vorticity equation (2.6a), by ignoring the mean-flow impact, we obtain the quasilinear balance between nonlinear wave term in (2.10c) and the external forcing, i.e. $\mathcal {N}_{W}\sim M_{F}k_{f}^{1/2}\mathsf {F}_1$, leading to $|\phi |^2\sim M_{F}k_{f}^{-1/2}f_0^{-2}$. The wave magnitude can be estimated by $|\phi |\sim RM_{F} m N^{-1} f_0^{-1}k_{f}^{-3/2}$ from the wave equation (2.6b). Combining them we obtain $R^2 \sim k_{f}^{3/2}N^2M_{F}^{-1}m^{-2}$. Since $\lambda \sim R^2$ and $\epsilon _{QG}\sim M_{F}^2$, the critical value scales as $\lambda _c \sim \epsilon _{QG}^{-1/2}N^2k_{f}f_0^{-1/2}m^{-2}$. Therefore in figure 9(b), we introduce a normalized parameter $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$ with $\epsilon _{QG}^0=8\times 10^{-5}$ to collapse data with different mean energy injection rates. Hence, we find that the quasilinear dynamics controls the phase transition of the NIW–QG coupled system.

Figure 10 shows the $\lambda$ dependence of the ratio between the upscale mean energy flux and mean energy injection rate, $\epsilon _{up}/\epsilon _{QG}$. Here, $\epsilon _{up} = \max _{K< k_{f}} |F_{QG}|$ is the upscale mean energy flux. In figure 10(a), we present the results with three mean energy injection rates. The upscale energy flux ratio gradually decreases as $\lambda$ increases and finally shows a second-order phase transition across $\lambda _c$, which contrasts with the behaviour of $\mathcal {E}_{jet}/\mathcal {E}_{QG}$ shown in figure 9, but these two quantities share the same critical values. This behaviour is similar to that on an $f$-plane (Xie Reference Xie2020). Figure 10(b) shows that after normalizing $\lambda$ by $\epsilon _{QG}^{-1/2}$, curves with different mean energy injection rates collapse. This implies that the above-mentioned quasilinear argument controls the phase transition of the NIW–QG coupled system. The concurrence of the disappearance of jet and upscale energy flux implies that we can understand jet annihilation from the perspective of NIW-modified mean energy flux. The injection of NIWs induced downscale flux of mean energy (Xie Reference Xie2020); therefore, the upscale energy flux decreases and the jets weaken. Thus, from the same perspective, other properties, such as jet spacing, should also be controlled by upscale energy flux, which we study in detail in § 5.

Figure 10. Ratio between the upscale energy flux and the total mean energy injection rate with different mean energy injection rates. In (b), the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

5. Jet spacing

Another important character of jets is their spacing, $L_{jet}$. We may define $L_{jet}$ from the jet energy spectrum:

(5.1)\begin{equation} L_{jet}=\frac{\displaystyle \int E_Z(k)\, {\rm d} k}{\displaystyle \int k E_Z(k)\, {\rm d} k}. \end{equation}

Proposed by Rhines (Reference Rhines1975), jet spacing can be captured by a characteristic scale $L_{Rh}=\sqrt {{U}/{\beta }}$, which is named the Rhines scale. By choosing the r.m.s. of jet velocity, $U_{jet} =\sqrt {2\mathcal {E}_{jet}}$, as the characteristic velocity, the Rhines scale can be expressed as

(5.2)\begin{equation} L_{Rh}=\frac{(2\mathcal{E}_{jet})^{1/4}}{\beta^{1/2}}. \end{equation}

In 2-D $\beta$-plane turbulence, energy transfers upscale, and with further frictional damping, the kinetic energy can be obtained from the balance between energy injection and dissipation as $\mathcal {E} = \epsilon /(2\alpha )$ in a statistically steady state. In our system, the QG mean energy transfers bidirectionally. Since the jet is a large-scale structure, one may propose that the upscale energy flux $\epsilon _{up}$ is important for jets (cf. Scott & Dritschel Reference Scott and Dritschel2012). Therefore, following the above-mentioned classic 2-D turbulence argument, we obtain a characteristic scale:

(5.3)\begin{equation} L_{\alpha}=\frac{\epsilon_{up}^{1/4}}{\alpha^{1/4}\beta^{1/2}}. \end{equation}

We further consider that in the NIW–QG coupled system, QG mean energy flux $F_{QG}$ can be decomposed into the mean part $F_{QG}^m$ and the NIW-related flux $F_{QG}^{NIW}$ (cf. (3.5)). Since a jet is a zonal-mean quantity, we conjecture that only $F_{QG}^m$ is relevant to the jet spacing. Therefore, we propose a characteristic scale estimated by the upscale energy flux induced by mean flow only, $\epsilon _{adv} = \max _{K< k_{f}} |F_{QG}^m|$:

(5.4)\begin{equation} L_{adv}=\frac{\epsilon_{adv}^{1/4}}{\alpha^{1/4}\beta^{1/2}}. \end{equation}

When no NIWs are injected into the coupled system (i.e. $\lambda =0$), ${L}_{adv}$ equals $L_{\alpha }$ since $\epsilon _{up} = \epsilon _{adv}$.

We show the normalized jet spacing and three normalized characteristic scales in figure 11(a) with $\lambda$ ranging from $0$ to $\lambda _c$. As $\lambda$ increases, jet spacing increases. However, both $L_{Rh}$ and $L_{\alpha }$ decrease, so they cannot describe jet spacing. Interestingly, even though the presence of NIWs reduces the upscale flux of the mean energy, the jet spacing increases with $\lambda$ increasing. In contrast, ${L}_{adv}$ increases as $\lambda$ increases, which can be seen from the increase of $F_{QG}^m$ (cf. figure 6). Figure 11(b) shows that the jet spacing is well described by ${L}_{adv}$, which implies that the upscale energy flux induced by mean advection controls jet formation. Thus, by measuring the mean quantities only, we can capture the jet spacing using the classic Rhines scale as if the NIW is absent, which again points out the catalytic role of NIWs.

Figure 11. (a) Jet spacings and different definitions of Rhines scale with $k_{f}=32$, $\epsilon _{QG}=8\times 10^{-5}$ and varying $\lambda$. The tilde denotes the scale normalized by NIW-free jet spacing ($\lambda =0$). (b) Comparison between the numerically obtained jet spacing $\tilde {L}_{jet}$ and the mean-advection-related Rhines scale $\tilde {L}_{adv}$ with different $\lambda$. In these simulations $\epsilon _{QG}=8\times 10^{-5}$.

6. Near-inertial-wave-induced symmetry breaking

The NIWs break the temporal symmetry by rotating in the direction opposite to that of the background rotation. This temporal symmetry breaking leads to spatial asymmetry, which we study in this section from the perspectives of mean-velocity profile and jet meridional migration.

6.1. Mean-velocity profile

In this subsection, we focus on the mean-velocity profile of zonal jets. We show the jet velocity profile $\bar {u}$ obtained by taking a zonal average of zonal velocity $u$ in figure 12(a). Similar to the classic $\beta$-plane jet profile, the zonal velocity has an east–west asymmetry: the east jets have cusps while the west jets are smooth. In figure 12(a), the presence of NIWs introduces an additional asymmetry of the jet with steeper left flanks (red rectangles) than the right flanks (blue rectangles). This asymmetry manifests itself in the asymmetric profile of zonal-mean vorticity $\bar {Q}=-\bar {u}_y$ with a greater magnitude of positive vorticity, which is shown in figure 12(b). Figure 12(c) shows that the zonal-mean wave action $|\bar {\phi }|^2$ concentrates around regions with negative vorticity, $\bar {Q} < 0$, which is consistent with NIW concentration on anticyclones (Kunze Reference Kunze1985; Danioux et al. Reference Danioux, Vanneste and Bühler2015). Since NIWs concentrate on the right flank with negative vorticity, the intensity of NIW–QG interaction there should be more potent than that at the left flank with positive vorticity. The NIW–QG interaction induces a downscale mean-energy flux and increases dissipation at small scales (cf. § 3). Therefore, the mean shear on the left flank is stronger than that on the right flank.

Figure 12. Zonal-mean profiles of (a) zonal velocity, (b) vorticity and (c) wave action with $\epsilon _{QG}=8\times 10^{-5}, \lambda =0.4$ at $t=2000$. The blue and red rectangles denote the regions with positive and negative vorticity, respectively.

To show the impact of NIW strength on mean-velocity profiles, in figure 13 we compare jet profiles with different $\lambda$ by moving the profiles to the same peak location. It shows that the asymmetry of jet flanks enhances as $\lambda$ increases. Since jets are annihilated when $\lambda >\lambda _c$, we only study the jet profile in the parameter regime with $\lambda <\lambda _c$.

Figure 13. Jet velocity profiles with $\epsilon _{QG}=8\times 10^{-5}$ and changing $\lambda$. Here, $\tilde {u}$ denotes the moved profiles sharing the same peak location for comparison; $\tilde {y}$ denotes the moved meridional coordinate.

To quantify the jet flank asymmetry, we study the skewness of the zonal-mean vorticity $\bar {Q}$. We define $Q_{3}=\langle \overline {Q}^3\rangle$ and $Q_{2}=\langle \overline {Q}^2\rangle$, and use the skewness $S_{k}=Q_{3}/Q_{2}^{3/2}$ to quantify the asymmetry. We present $S_{k}$ with different $\epsilon _{QG}$ and $\lambda$ in figure 14(a). The skewness of the mean vorticity first increases with $\lambda$ due to NIWs concentrating on the negative vorticity, and then it suddenly drops as $\lambda$ reaches $\lambda _c$ because jets are annihilated. Similar to figures 9(b) and 10(b), figure 14(b) shows that the same normalization $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$ also works for $S_{k}$, which again reveals the significance of quasilinear dynamics (cf. § 4).

Figure 14. The skewness of zonal-mean vorticity with different mean energy injection rates. In (b) the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

Taken together, these results suggest that the mean flow on the right flank is reduced more than that on the left flank due to NIWs concentrating on the right flanks, where anticyclones reside and NIW–mean interaction grows. Therefore, NIWs catalyse the asymmetry of jet flanks. Besides, in the current parameter regime, we find that the Stokes drift (Bühler Reference Bühler2009) is negligible compared with the Lagrangian mean velocity. Details are presented in Appendix A.

6.2. Jet migration

In this subsection, we focus on the impact of NIWs on jet migration. We show the Hovmöller diagram of the zonal-mean zonal velocity, $\bar {u}$, with $\lambda =0.4$ and $\epsilon _{QG}=8\times 10^{-5}$ in figure 15, where jets migrate with a constant speed in the meridional direction. From this Hovmöller diagram, we can define a jet migration speed $\mathcal {V}=\delta y/\delta t$, where $\delta y$ is the latitudinal shift and $\delta t$ is the elapsed time.

Figure 15. Hovmöller diagram of the zonal-mean zonal velocity with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

Figure 16 shows the dependence of the measured jet migration speed on the energy injection ratio. The zonal jets without NIWs stay static, and as the NIW energy injection rate increases, jet meridional migration speed gradually increases. To find the mechanism behind zonal jet migration, we start from the zonal-mean momentum equation derived from (2.6a):

(6.1)\begin{equation} \partial_t \bar{u} +\partial_y \overline{u'v'} -\bar{\varLambda} ={-}\alpha_1 \bar{u}+\nu_1 \frac{\partial^6\bar{u}}{\partial y^{6}},\end{equation}

where $(u',v')$ is the fluctuation with respect to the zonal-mean velocity (cf. Chemke & Kaspi Reference Chemke and Kaspi2015b). Parameter $\varLambda$ is derived from the wave-related term in (2.10c) with $\hat {\varLambda }=-\mathrm {i}\hat {\mathcal {N}}_{NIW}/l$ in Fourier space. Using Taylor's identity

(6.2)\begin{equation} -\partial_y ( \overline{u'v'})=\overline{v'Q'}, \end{equation}

equation (6.1) becomes

(6.3)\begin{equation} \partial_t \bar{u} =\overline{v'Q'} + \bar{\varLambda} -\alpha_1 \bar{u} +\nu_1 \frac{\partial^6 \bar{u}}{\partial y^{6}}.\end{equation}

Here, $Q'= Q-\bar {Q}$. Assuming that jets migrate in the meridional direction with a constant speed, we consider a frame of reference moving with jets by introducing a coordinate $\varphi =y-\mathcal {V}^*t$ with $\mathcal {V}^*$ the jet migration speed. Thus (6.3) becomes

(6.4)\begin{equation} -\mathcal{V}^*\frac{\mathrm{d} \bar{u}}{\mathrm{d} \varphi}= \underset{\mathcal{T}(\varphi)}{\underbrace{ \overline{v'Q'} + \bar{\varLambda} -\alpha_1 \bar{u} +\nu_1 \frac{\partial^6 \bar{u}}{\partial y^{6}}}},\end{equation}

where we introduce the momentum flux convergence $\mathcal {T}$ for convenience.

Figure 16. Dependence of jet migration speed $\mathcal {V}$ on $\lambda$ with $\epsilon _{QG}=8\times 10^{-5}$.

In figure 17, we show the NIW-induced asymmetric jet profile (cf. § 6.1) and its corresponding momentum flux convergence $\mathcal {T}$. We define the width of the jet flank as the meridional range with $\bar {u}>0$, i.e. $\varphi _0^+-\varphi _0^-$, and $\varphi _{jet}$ is the latitude of the maximum zonal-mean zonal velocity $\bar {u}_{max}$. Integrating (6.4) from $\varphi _0^-$ to $\varphi _{jet}$ and $\varphi _{jet}$ to $\varphi _0^+$ and then subtracting the resultant equations, the jet migration speed $\mathcal {V}^*$ is expressed as

(6.5)\begin{equation} \mathcal{V}^*=\frac{1}{2\bar{u}_{max}} \left( \int^{\varphi_0^+}_{\varphi_{jet}} \mathcal{T}\,\mathrm{d} \varphi - \int^{\varphi_{jet}}_{\varphi_0^-} \mathcal{T}\,\mathrm{d} \varphi \right).\end{equation}

Figure 17. Profiles of the normalized zonal-mean velocity $\bar {u}/\bar {u}_{max}$ and the normalized momentum flux convergence $\mathcal {T}/\mathcal {T}_{max}$ with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

To find the dominant effect on jet migration, we introduce the following decomposition:

(6.6)\begin{equation} \int^{\varphi_0^+}_{\varphi_{jet}}\mathcal{T}\,\mathrm{d} \varphi -\int^{\varphi_{jet}}_{\varphi_0^-}\mathcal{T}\,\mathrm{d} \varphi =\mathcal{I}_{adv}+\mathcal{I}_{NIW} +\mathcal{I}_{\alpha}+\mathcal{I}_{\nu},\end{equation}

where

(6.7a)$$\begin{gather} \mathcal{I}_{adv}=\int^{\varphi_0^+}_{\varphi_{jet}}\overline{v'Q'}\,\mathrm{d} \varphi - \int^{\varphi_{jet}}_{\varphi_0^-}\overline{v'Q'}\,\mathrm{d} \varphi, \end{gather}$$
(6.7b)$$\begin{gather}\mathcal{I}_{NIW}=\int^{\varphi_0^+}_{\varphi_{jet}}\bar{\varLambda}\,\mathrm{d} \varphi - \int^{\varphi_{jet}}_{\varphi_0^-}\bar{\varLambda}\,\mathrm{d} \varphi, \end{gather}$$
(6.7c)$$\begin{gather}\mathcal{I}_{\alpha}=\int^{\varphi_0^+}_{\varphi_{jet}}-\alpha_1 \bar{u}\,\mathrm{d} \varphi + \int^{\varphi_{jet}}_{\varphi_0^-}\alpha_1 \bar{u}\,\mathrm{d} \varphi, \end{gather}$$
(6.7d)$$\begin{gather}\mathcal{I}_{\nu}=\int^{\varphi_0^+}_{\varphi_{jet}}\nu_1 \frac{\partial^6\bar{u}}{\partial \varphi^{6}} \,\mathrm{d} \varphi - \int^{\varphi_{jet}}_{\varphi_0^-}\nu_1 \frac{\partial^6\bar{u}}{\partial \varphi^{6}} \,\mathrm{d} \varphi. \end{gather}$$

Figure 18(a) shows the $\lambda$ dependence of these terms. We observe that the total net momentum flux, $\mathcal {I}_{total}=\mathcal {I}_{adv}+\mathcal {I}_{NIW}+ \mathcal {I}_{\alpha }+ \mathcal {I}_{\nu }$, increases with $\lambda$. Therein, the dissipation-related net momentum fluxes, $\mathcal {I}_{\alpha }$ and $\mathcal {I}_{\nu }$, are negligible. The advection-related net momentum flux, $\mathcal {I}_{adv}$, is dominant and increases with $\lambda$, while the wave-related net momentum flux, $\mathcal {I}_{NIW}$, has an opposite sign to $\mathcal {I}_{total}$ and therefore prohibits jet migration. Figure 18(b) shows that when $\lambda$ ranges from $0$ to $\lambda _c$, the jet migration speeds calculated from (6.5) capture the migration speeds obtained from numerical simulations. The impact of NIWs on jet migration again shows its catalytic effect – the direct effect of NIWs, $\mathcal {I}_{NIW}$, restrains jet migration, but NIW-modified mean flow dominates and drives the migration.

Figure 18. (a) The dependences of the net momentum flux $\mathcal {I}$ on $\lambda$. (b) Numerically obtained jet migration speed $\mathcal {V}$ against the migration speed $\mathcal {V}^*$ calculated from momentum balance (6.5). Here, $\epsilon _{QG}=8\times 10^{-5}$.

The direction of jet migration can be understood as a consequence of NIW concentration in anticyclones (Danioux et al. Reference Danioux, Vanneste and Bühler2015). Since NIWs accumulate in anticyclones, more energy is transferred downscale by negative vorticity and finally dissipates (cf. § 3), so the negative vorticity of the mean flow is weaker than the positive vorticity. Then the vortex pairs propagate poleward. This phenomenon was observed in simulations of both primitive equations (Hernandez-Duenas, Smith & Stechmann Reference Hernandez-Duenas, Smith and Stechmann2014) and reduced model (Xie & Vanneste Reference Xie and Vanneste2015). Similarly, figure 19 shows that both jets and waves migrate poleward and NIWs concentrate in regions with negative vorticity, so jets migrate poleward.

Figure 19. Hovmöller diagram of the zonal-mean (a) vorticity and (b) NIW amplitude with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

We need to note that the coupled model was initially derived in the ocean context. But most jets in ocean models are found to propagate equatorward (Chan et al. Reference Chan, Plumb and Cerovecki2007; Ashkenazy & Tziperman Reference Ashkenazy and Tziperman2016). Nevertheless, poleward migration was also observed over a short range and less systematically (Chan et al. Reference Chan, Plumb and Cerovecki2007). Thus, the impact of NIWs cannot fully explain the migration of oceanic jets.

7. Summary and discussion

Using a 2-D single-vertical-mode NIW–QG coupled model on a $\beta$-plane (2.1), we study the impact of NIWs on zonal jets in forced-dissipative statistically steady states. We explore the parameter regime with different mean energy injection rates and the ratio between NIW and QG mean energy injection rate, which $\lambda$ parametrizes.

The presence of NIWs leads to bidirectional mean-flow energy transfer, which shows a second-order phase transition with a critical value $\lambda _c$. When $0<\lambda <\lambda _c$, the energy of balanced flow transfers to both large and small scales. As $\lambda$ increases, the portion of upscale transferring energy decreases, while the proportion of zonal-jet energy in the energy of the balanced flow remains almost unchanged. When $\lambda > \lambda _c$, the QG inverse cascade is eliminated and zonal jets cannot form. Since there is a negligible energy exchange between mean flows and NIWs, the existence of the catalytic wave induction mechanism initially proposed on an $f$-plane (Xie Reference Xie2020) is justified on a $\beta$-plane. This mechanism differs from the spontaneous and stimulated loss of balance (Vanneste Reference Vanneste2013; Xie & Vanneste Reference Xie and Vanneste2015), where the mean energy converts to that of NIWs and finally dissipates.

Based on the picture that the $\beta$-effect arrests the upscale energy flux, previous studies (Rhines Reference Rhines1975; Williams Reference Williams1978) showed that the Rhines scale could capture the zonal jet spacing. Nevertheless, we find that the Rhines scale calculated from the r.m.s. velocity (Rhines Reference Rhines1975) and the upscale energy transfer rate $\epsilon _{up}$ (Vallis Reference Vallis2006) cannot sufficiently describe jet spacing when NIWs are present. Instead, the Rhines scale calculated from the upscale-mean-energy transfer rate induced by the mean advection, $L_{adv}=(2\epsilon _{adv})^{1/4}\alpha ^{-1/4}\beta ^{-1/2}$, captures the jet spacing before jet total annihilation by NIWs.

The concentration of NIWs in anticyclones induces symmetry breaking of zonal jets. The NIWs lead to an asymmetry of the left and right flanks about the jet profile cusp, where the right flank's slope is flatter than that of the left flank. Moreover, this asymmetry enhances as the energy injection ratio $\lambda$ increases. The concentration of NIWs in anticyclones or regions with negative vorticity (Danioux et al. Reference Danioux, Vanneste and Bühler2015) can explain the mechanism behind such asymmetry. Therefore at the right flank of a jet cusp, stronger NIW–mean flow interaction leads to stronger downscale mean energy flux induced by waves, which enhances the mean flow dissipation, so the right flank is flatter than the left flank. The NIW-induced symmetry breaking also manifests in jets’ poleward migration. The interaction between NIWs and mean flows also influences the jet migration speed in the meridional direction. Based on an expression of migration speed calculated from the momentum equation in a frame of reference moving with jets (Chemke & Kaspi Reference Chemke and Kaspi2015b), we find that the distribution of the momentum flux convergence to the cusps in the velocity profile is asymmetric, which leads to jet migration. In this expression, the mean-advection net momentum flux dominates, while the NIW-related term impedes jet migration but is of secondary importance. Furthermore, we numerically find that the migration speed increases as the energy injection of NIWs increases. Even though the asymmetric jet profile and meridional migration coexist, we do not know if this coexistence is inevitable. Such coexistence is also observed in jet migration without wave impact (cf. Cope Reference Cope2020), so there is more to be understood on jet asymmetry.

Considering that the strong vortex regions are close to the peak of the cusp of eastward jets, this poleward migration mimics the left turning of vortex dipoles (Hernandez-Duenas et al. Reference Hernandez-Duenas, Smith and Stechmann2014; Xie & Vanneste Reference Xie and Vanneste2015). Even though the reduced model is derived in an oceanic context (Xie & Vanneste Reference Xie and Vanneste2015), the jet migration direction coincides with the results of observations (Riehl et al. Reference Riehl, Yeh and La Seur1950; Dickey et al. Reference Dickey, Marcus and Hide1992) and numerical simulations (James & Dodd Reference James and Dodd1996; Chemke & Kaspi Reference Chemke and Kaspi2015b) in the atmosphere. However, in the ocean, poleward migration of zonal flow is observed less frequently (Chan et al. Reference Chan, Plumb and Cerovecki2007), implying that the impact of NIWs on zonal jets may not be a dominant effect that determines the direction of oceanic jet migration.

The catalytic wave impact on zonal jets manifests itself from three aspects. (i) In the mean-flow bidirectional energy transfer across scales, NIWs impact the relative strength of the downscale flux without energy conversion between NIW and mean flow. (ii) The jet spacing is described by the Rhines scale obtained from the upscale mean energy flux induced by mean flow alone as if there is no NIW. (iii) The jet meridional migration speed is dominated by the asymmetry of mean momentum flux induced by mean flow, but in our set-up, this asymmetry is induced by NIWs. We summarize the effects of NIW concentration on the right flank, the asymmetry of the jet and the meridional migration in figure 20. However, we do not understand why these wave effects are catalytic, which is an interesting future direction.

Figure 20. Illustration of the jet–NIW interaction.

We close this paper by pointing out several limitations of the present study. To solve the coupled NIW–QG model with the pseudospectral method, we prescribe a periodic $\beta (y)$ (cf. 2.4). Therefore, around one-half of the meridional extension approximates a $\beta$-plane, while other regions are only introduced to construct a periodic domain for the spectral method. To avoid the Gibbs effect, the $\beta$-effect is artificially prescribed with a narrow but finite range of non-constant value, which leads to unphysical effects, but we assume that this effect is only confined in this narrow region, which is the drawback of this set-up. In addition, the NIW–QG model in this paper is asymptotically derived in the regime of small Rossby and small Burger numbers (cf. Xie & Vanneste Reference Xie and Vanneste2015), which limits the validity of the current results; therefore, it would be interesting to check if our results persist in primitive equations in future work. Besides, we applied a simple turbulence phenomenology based on energy flux to explain jet formation, but we did not distinguish the local and non-local effects in spectral space. So in future work, we plan to study the impact of waves on jets by applying a quasilinear approximation to the balanced motion (cf. Marston & Tobias Reference Marston and Tobias2023).

Funding

The authors gratefully acknowledge financial support from the National Natural Science Foundation of China (NSFC) under grant nos. 92052102 and 12272006 and Joint Laboratory of Marine Hydrodynamics and Ocean Engineering, Pilot National Laboratory for Marine Science and Technology (Qingdao) under grant no. 2022QNLM010201.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Stokes drift

The vertical plane-wave displacement and velocity are expressed by the complex amplitude $\phi$ as

(A1a,b)\begin{equation} \xi+\mathrm{i} \eta=\phi\, \mathrm{e}^{-\mathrm{i} f_0 t+ \mathrm{i} m z}, \quad \zeta=\frac{\mathrm{i}}{m}\phi_s\, \mathrm{e}^{-\mathrm{i} f_0 t+ \mathrm{i} m z} +{\rm c.c.} \end{equation}

and

(A2a,b)\begin{equation} u_0+\mathrm{i} v_0 ={-}\mathrm{i} f_0 \phi \, \mathrm{e}^{-\mathrm{i} f_0 t+\mathrm{i} m z}, \quad w_0 =\frac{f_0}{m}\phi_s \, \mathrm{e}^{-\mathrm{i} f_0 t+\mathrm{i} m z} +{\rm c.c.}, \end{equation}

where $s=x+\mathrm {i} y,\partial _s=(\partial _x-\mathrm {i} \partial _y)/2$ and ‘c.c.’ denotes the complex conjugate. The Lagrangian mean velocity differs from the Eulerian mean velocity by a Stokes drift (Bühler Reference Bühler2009), which is expressed as

(A3)\begin{equation} \bar{\boldsymbol{u}}^S=\overline{\boldsymbol{\xi}\boldsymbol{\cdot} \boldsymbol{\nabla} \boldsymbol{u}_0}. \end{equation}

To check the difference between the Lagrangian and Eulerian mean velocity, we calculate the Stokes drift by substituting (A1a,b) and (A2a,b) into (A3):

(A4)\begin{equation} \overline{\boldsymbol{u}}^S=\left(-\frac{f_0}{2}\nabla^\perp \vert\phi\vert^2,\frac{f_0}{4m}(\phi^*{\rm \Delta} \phi+{\rm c.c.}) +\frac{f_0}{2m}[\vert\phi_y\vert^2+\vert\phi_x\vert^2+\mathrm{i} (\phi_x\phi_y^*-{\rm c.c.})]\right). \end{equation}

Here, wave action acts as the streamfunction of the horizontal components of Stokes drift (Rocha et al. Reference Rocha, Wagner and Young2018). Since NIWs concentrate in regions with negative vorticity (Danioux et al. Reference Danioux, Vanneste and Bühler2015), the perpendicular gradient of wave action is large at the peak of the mean velocity profile. Therefore, Stokes drift peaks around the maximum Lagrangian mean velocity. In figure 21, the Stokes drift is much weaker than the Lagrangian mean flow, so in the main text we do not pay special attention to the Eulerian mean velocity.

Figure 21. The Lagrangian mean velocity profile $\bar {u}$ and the Stokes drift $\bar {u}^S$ with $\epsilon _{QG}=8\times 10^{-5}$ and $R=2.0$.

References

Abernathey, R., Marshall, J. & Ferreira, D. 2011 The dependence of southern ocean meridional overturning on wind stress. J. Phys. Oceanogr. 41 (12), 22612278.CrossRefGoogle Scholar
Andrews, D.G. & Mcintyre, M.E. 1978 An exact theory of nonlinear waves on a Lagrangian-mean flow. J. Fluid Mech. 89 (4), 609646.CrossRefGoogle Scholar
Ashkenazy, Y. & Tziperman, E. 2016 Variability, instabilities, and eddies in a snowball ocean. J. Clim. 29 (2), 869888.CrossRefGoogle Scholar
Asselin, O., Thomas, L.N., Young, W.R. & Rainville, L. 2020 Refraction and straining of near-inertial waves by barotropic eddies. J. Phys. Oceanogr. 50, 34393454.CrossRefGoogle Scholar
Asselin, O. & Young, W.R. 2019 An improved model of near-inertial wave dynamics. J. Fluid Mech. 876, 428448.CrossRefGoogle Scholar
Asselin, O. & Young, W.R. 2020 Penetration of wind-generated near-inertial waves into a turbulent ocean. J. Phys. Oceanogr. 50 (6), 16991716.CrossRefGoogle Scholar
Barkan, R., Srinivasan, K., Yang, L., McWilliams, J.C., Gula, J. & Vic, C. 2021 Oceanic mesoscale eddy depletion catalyzed by internal waves. Geophys. Res. Lett. 48, e2021GL094376.CrossRefGoogle Scholar
Barkan, R., Winters, K.B. & McWilliams, J.C. 2017 Stimulated imbalance and the enhancement of eddy kinetic energy dissipation by internal waves. J. Phys. Oceanogr. 47 (1), 181198.CrossRefGoogle Scholar
Berloff, P., Kamenkovich, I. & Pedlosky, J. 2009 A mechanism of formation of multiple zonal jets in the oceans. J. Fluid Mech. 628, 395425.CrossRefGoogle Scholar
Bouchet, F., Marston, J.B. & Tangarife, T. 2018 Fluctuations and large deviations of reynolds stresses in zonal jet dynamics. Phys. Fluids 30 (1), 015110.CrossRefGoogle Scholar
Bouchet, F., Rolland, J. & Simonnet, E. 2019 Rare event algorithm links transitions in turbulent flows with activated nucleations. Phys. Rev. Lett. 122 (7), 074502.CrossRefGoogle ScholarPubMed
Bouchet, F. & Venaille, A. 2012 Statistical mechanics of two-dimensional and geophysical flows. Phys. Rep. 515 (5), 227295.CrossRefGoogle Scholar
Bühler, O. 2009 Waves and Mean Flows. Cambridge University Press.CrossRefGoogle Scholar
Chan, C.J., Plumb, R.A. & Cerovecki, I. 2007 Annular modes in a multiple migrating zonal jet regime. J. Atmos. Sci. 64 (11), 40534068.CrossRefGoogle Scholar
Chekhlov, A., Orszag, S.A., Sukoriansky, S., Galperin, B. & Staroselsky, I. 1996 The effect of small-scale forcing on large-scale structures in two-dimensional flows. Phys. D 98 (2-4), 321334.CrossRefGoogle Scholar
Chemke, R. & Kaspi, Y. 2015 a The latitudinal dependence of atmospheric jet scales and macroturbulent energy cascades. J. Atmos. Sci. 72 (10), 38913907.CrossRefGoogle Scholar
Chemke, R. & Kaspi, Y. 2015 b Poleward migration of eddy-driven jets: poleward migration of eddy-driven jets. J. Adv. Model. Earth Syst. 7 (3), 14571471.CrossRefGoogle Scholar
Cope, L. 2020 The dynamics of geophysical and astrophysical turbulence. PhD Thesis, University of Cambridge.Google Scholar
Cox, S.M. & Matthews, P.C. 2002 Exponential time differencing for stiff systems. J. Comput. Phys. 176 (2), 430455.CrossRefGoogle Scholar
Danilov, S.D. & Gurarie, D 2000 Quasi-two-dimensional turbulence. Phys. Uspekhi 43 (9), 863900.CrossRefGoogle Scholar
Danilov, S. & Gurarie, D. 2004 Scaling, spectra and zonal jets in beta-plane turbulence. Phys. Fluids 16 (7), 2592–2603.CrossRefGoogle Scholar
Danioux, E., Vanneste, J. & Bühler, O. 2015 On the concentration of near-inertial waves in anticyclones. J. Fluid Mech. 773, R2.CrossRefGoogle Scholar
Danioux, E., Vanneste, J., Klein, P. & Sasaki, H. 2012 Spontaneous inertia-gravity-wave generation by surface-intensified turbulence. J. Fluid Mech. 699, 153173.CrossRefGoogle Scholar
Dickey, J.O., Marcus, S.L. & Hide, R. 1992 Global propagation of interannual fluctuations in atmospheric angular momentum. Nature 357 (6378), 484488.CrossRefGoogle Scholar
Dritschel, D.G. & McIntyre, M.E. 2008 Multiple jets as PV staircases: the phillips effect and the resilience of eddy-transport barriers. J. Atmos. Sci. 65 (3), 855874.CrossRefGoogle Scholar
Dunkerton, T.J. & Scott, R.K. 2008 A barotropic model of the angular momentum–conserving potential vorticity staircase in spherical geometry. J. Atmos. Sci. 65 (4), 11051136.CrossRefGoogle Scholar
Elipot, S., Lumpkin, R. & Prieto, G. 2010 Modification of inertial oscillations by the mesoscale eddy field. J. Geophys. Res. 115, C09010.Google Scholar
Farrell, B.F. & Ioannou, P.J. 2003 Structural stability of turbulent jets. J. Atmos. Sci. 60 (17), 21012118.2.0.CO;2>CrossRefGoogle Scholar
Feldstein, S.B. 1998 An observational study of the intraseasonal poleward propagation of zonal mean flow anomalies. J. Atmos. Sci. 55 (15), 25162529.2.0.CO;2>CrossRefGoogle Scholar
Ferrari, R. & Wunsch, C. 2009 Ocean circulation kinetic energy: reservoirs, sources, and sinks. Annu. Rev. Fluid Mech. 41 (1), 253282.CrossRefGoogle Scholar
Fu, L.-L. 1981 Observations and models of inertial waves in the deep ocean. Rev. Geophys. Space Phys. 19 (1), 141170.CrossRefGoogle Scholar
Galperin, B. & Read, P.L. (Eds.) 2019 Zonal Jets: Phenomenology, Genesis, and Physics, 1st edn. Cambridge University Press.CrossRefGoogle Scholar
Galperin, B., Sukoriansky, S. & Dikovskaya, N. 2010 Geophysical flows with anisotropic turbulence and dispersive waves: flows with a $\beta$-effect. Ocean Dyn. 60 (2), 427441.CrossRefGoogle Scholar
Galperin, B., Sukoriansky, S., Dikovskaya, N., Read, P.L., Yamazaki, Y.H. & Wordsworth, R. 2006 Anisotropic turbulence and zonal jets in rotating flows with a $\beta$-effect. Nonlinear Process. Geophys. 13 (1), 8398.CrossRefGoogle Scholar
García-Melendo, E. 2001 A study of the stability of Jovian zonal winds from HST images: 1995–2000. Icarus 152 (2), 316330.CrossRefGoogle Scholar
Gertz, A. & Straub, D.N. 2009 Near-inertial oscillations and the damping of midlatitude gyres: a modeling study. J. Phys. Oceanogr. 39 (9), 23382350.CrossRefGoogle Scholar
Grisouard, N. & Thomas, L.N. 2016 Energy exchanges between density fronts and near-inertial waves reflecting off the ocean surface. J. Phys. Oceanogr. 46 (2), 501516.CrossRefGoogle Scholar
Hernandez-Duenas, G., Smith, L.M. & Stechmann, S.N. 2014 Investigation of Boussinesq dynamics using intermediate models based on wave–vortical interactions. J. Fluid Mech. 747, 247287.CrossRefGoogle Scholar
James, I.N. & Dodd, J.P. 1996 A mechanism for the low-frequency variability of the mid-latitude troposphere. Q. J. R. Meteorol. Soc. 122 (533), 11971210.CrossRefGoogle Scholar
Joyce, T.M., Toole, J.M., Klein, P. & Thomas, L.N. 2013 A near-inertial mode observed within a gulf stream warm-core ring. J. Geophys. Res. Oceans 118 (4), 17971806.CrossRefGoogle Scholar
Kafiabad, H.A., Vanneste, J. & Young, W.R. 2021 Interaction of near-inertial waves with an anticyclonic vortex. J. Phys. Oceanogr. 51, 20352048.Google Scholar
Kraichnan, R.H. 1967 Inertial ranges in two-dimensional turbulence. Phys. Fluids 10 (7), 14171423.CrossRefGoogle Scholar
Kunze, E. 1985 Near-inertial wave propagation in geostrophic shear. J. Phys. Oceanogr. 15 (5), 544565.2.0.CO;2>CrossRefGoogle Scholar
Kunze, E., Schmitt, R.W. & Toole, J.M. 1995 The energy balance in a warm-core ring's near-inertial critical layer. J. Phys. Oceanogr. 25 (5), 942957.2.0.CO;2>CrossRefGoogle Scholar
Lee, D.-K. & Niiler, P.P. 1998 The inertial chimney: the near-inertial energy drainage from the ocean surface to the deep layer. J. Geophys. Res.: Oceans 103 (C4), 75797591.CrossRefGoogle Scholar
Lemasquerier, D., Favier, B. & Le Bars, M. 2021 Zonal jets at the laboratory scale: hysteresis and Rossby waves resonance. J. Fluid Mech. 910, A18.CrossRefGoogle Scholar
Lighthill, J. 1978 Waves in Fluids. Cambridge University Press.Google Scholar
Maltrud, M.E. & Vallis, G.K. 1991 Energy spectra and coherent structures in forced two-dimensional and beta-plane turbulence. J. Fluid Mech. 228, 321342.Google Scholar
Manfroi, A.J. & Young, W.R. 1999 Slow evolution of zonal jets on the beta plane. J. Atmos. Sci. 56 (5), 784800.2.0.CO;2>CrossRefGoogle Scholar
Marcus, P.S. & Lee, C. 1998 A model for eastward and westward jets in laboratory experiments and planetary atmospheres. Phys. Fluids 10 (6), 14741489.CrossRefGoogle Scholar
Marston, J.B. & Tobias, S.M. 2023 Recent developments in theories of inhomogeneous and anisotropic turbulence. Annu. Rev. Fluid Mech. 55 (1), 351375.CrossRefGoogle Scholar
Nagai, T., Tandon, A., Kunze, E. & Mahadevan, A. 2015 Spontaneous generation of near-inertial waves by the kuroshio front. J. Phys. Oceanogr. 45 (9), 23812406.CrossRefGoogle Scholar
Okuno, A. & Masuda, A. 2003 Effect of horizontal divergence on the geostrophic turbulence on a beta-plane: suppression of the Rhines effect. Phys. Fluids 15 (1), 5665.CrossRefGoogle Scholar
Plougonven, R. & Snyder, C. 2007 Inertia–gravity waves spontaneously generated by jets and fronts. Part I: different baroclinic life cycles. J. Atmos. Sci. 64, 25022520.CrossRefGoogle Scholar
Rhines, P.B. 1975 Waves and turbulence on a beta-plane. J. Fluid Mech. 69 (3), 417443.CrossRefGoogle Scholar
Rhines, P.B. 1979 Geostrophic turbulence. Annu. Rev. Fluid Mech. 11, 401441.CrossRefGoogle Scholar
Rhines, P.B. 1994 Jets. Chaos 4 (2), 313339.CrossRefGoogle ScholarPubMed
Riehl, H., Yeh, T.C. & La Seur, N.E. 1950 A study of variations of the general circulation. J. Meteorol. 7 (3), 181194.2.0.CO;2>CrossRefGoogle Scholar
Rocha, C.B., Wagner, G.L. & Young, W.R. 2018 Stimulated generation: extraction of energy from balanced flow by near-inertial waves. J. Fluid Mech. 847, 417451.CrossRefGoogle Scholar
Rolland, J. & Simonnet, E. 2015 Statistical behaviour of adaptive multilevel splitting algorithms in simple models. J. Comput. Phys. 283, 541558.CrossRefGoogle Scholar
Salmon, R. 2013 An alternative view of generalized Lagrangian mean theory. J. Fluid Mech. 719, 165182.CrossRefGoogle Scholar
Sanchez-Lavega, A. & Rojas, J.F. 2000 Saturn's zonal winds at cloud level. Icarus 147 (2), 405420.CrossRefGoogle Scholar
Sánchez-Lavega, A., et al. 2008 Depth of a strong jovian jet from a planetary-scale disturbance driven by storms. Nature 451 (7177), 437440.CrossRefGoogle Scholar
Scott, R.K. & Dritschel, D.G. 2012 The structure of zonal jets in geostrophic turbulence. J. Fluid Mech. 711, 576598.CrossRefGoogle Scholar
Shakespeare, C.J. & McC. Hogg, A. 2018 The life cycle of spontaneously generated internal waves. J. Phys. Oceanogr. 48 (2), 343359.CrossRefGoogle Scholar
Smith, K.S. 2004 A local model for planetary atmospheres forced by small-scale convection. J. Atmos. Sci. 61 (12), 14201433.2.0.CO;2>CrossRefGoogle Scholar
Smith, L.M. & Waleffe, F. 2002 Generation of slow large scales in forced rotating stratified turbulence. J. Fluid Mech. 451, 145168.CrossRefGoogle Scholar
Soward, A.M. & Roberts, P.H. 2010 The hybrid Euler–Lagrange procedure using an extension of Moffatt's method. J. Fluid Mech. 661, 4572.CrossRefGoogle Scholar
Srinivasan, K. & Young, W.R. 2012 Zonostrophic instability. J. Atmos. Sci. 69 (5), 16331656.CrossRefGoogle Scholar
Sukoriansky, S., Dikovskaya, N. & Galperin, B. 2007 On the arrest of inverse energy cascade and the rhines scale. J. Atmos. Sci. 64 (9), 33123327.CrossRefGoogle Scholar
Taylor, S. & Straub, D. 2016 Forced near-inertial motion and dissipation of ow-frequency kinetic energy in a wind-driven channel flow. J. Phys. Oceanogr. 46 (1), 7993.CrossRefGoogle Scholar
Thomas, L.N. 2012 On the effects of frontogenetic strain on symmetric instability and inertia–gravity waves. J. Fluid Mech. 711, 620640.CrossRefGoogle Scholar
Thomas, J. & Arun, S. 2020 Near-inertial waves and geostrophic turbulence. Phys. Rev. Fluid 5 (1), 014801.CrossRefGoogle Scholar
Thomas, L.N., Rainville, L., Asselin, O., Young, W.R., Girton, J., Whalen, C.B., Centurioni, L. & Hormann, V. 2020 Direct observations of near-inertial wave $\zeta$ -refraction in a dipole vortex. Geophys. Res. Lett. 47, e2020GL090375.CrossRefGoogle Scholar
Uppala, S.M., et al. 2005 The ERA-40 re-analysis. J. Atmos. Sci. 131 (612), 29613012.Google Scholar
Vallis, G.K. 2006 Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press.CrossRefGoogle Scholar
Vallis, G.K. & Maltrud, M. 1993 Generation of mean flows and jets on a beta plane and over topography. J. Phys. Oceanogr. 23, 13461362.2.0.CO;2>CrossRefGoogle Scholar
Vanneste, J. 2008 Exponential smallness of inertia–gravity wave generation at small Rossby number. J. Atmos. Sci. 65, 16221637.CrossRefGoogle Scholar
Vanneste, J. 2013 Balance and spontaneous wave generation in geophysical flows. Annu. Rev. Fluid Mech. 45 (1), 147172.CrossRefGoogle Scholar
Vasavada, A.R. & Showman, A.P. 2005 Jovian atmospheric dynamics: an update after Galileo and Cassini. Rep. Prog. Phys. 68 (8), 19351996.CrossRefGoogle Scholar
Wagner, G.L. & Young, W.R. 2016 A three-component model for the coupled evolution of near-inertial waves, quasi-geostrophic flow and the near-inertial second harmonic. J. Fluid Mech. 802, 806837.CrossRefGoogle Scholar
Whitt, D.B. & Thomas, L.N. 2015 Resonant generation and energetics of wind-forced near-inertial motions in a geostrophic flow. J. Phys. Oceanogr. 45, 181208.CrossRefGoogle Scholar
Williams, G.P. 1978 Planetary circulations: 1. Barotropic representation of jovian and terrestrial turbulence. J. Atmos. Sci. 35 (8), 13991426.2.0.CO;2>CrossRefGoogle Scholar
Williams, G.P. 2003 Jovian dynamics. Part III: multiple, migrating, and equatorial jets. J. Atmos. Sci. 60 (10), 12701296.2.0.CO;2>CrossRefGoogle Scholar
Woillez, E. & Bouchet, F. 2019 Barotropic theory for the velocity profile of Jupiter's turbulent jets: an example for an exact turbulent closure. J. Fluid Mech. 860, 577607.CrossRefGoogle Scholar
Xie, J.-H. 2020 Downscale transfer of quasigeostrophic energy catalyzed by near-inertial waves. J. Fluid Mech. 904, A40.CrossRefGoogle Scholar
Xie, J.-H. & Vanneste, J. 2015 A generalised-Lagrangian-mean model of the interactions between near-inertial waves and mean flow. J. Fluid Mech. 774, 143169.CrossRefGoogle Scholar
Young, W. & Ben Jelloul, M. 1997 Propagation of near-inertial oscillations through a geostrophic flow. J. Mar. Res. 55, 735766.CrossRefGoogle Scholar
Zeitlin, V. 2008 Decoupling of balanced and unbalanced motions and inertia–gravity wave emission: small versus large Rossby numbers. J. Atmos. Sci. 65, 3528–354.CrossRefGoogle Scholar
Figure 0

Figure 1. Illustration of (a) $\beta (y)$ and (b) $g(y)$ with $L_y=2{\rm \pi}$ and $l_b=10$ in (2.4).

Figure 1

Figure 2. A snapshot of QG mean flow vorticity fields in the statistically steady state of the NIW-free simulation with $\epsilon _{QG}=8\times 10^{-5}$.

Figure 2

Figure 3. Evolutions of (a) QG mean, (b) $\beta$-related, (c) $\mathrm {NIW}$ and (d) total energy (cf. (2.5a)) in simulations with $\lambda =0.016, 0.144, 0.4, 0.784, 1.296$. In (b), for clarity we vertically shift the curves following $\mathcal {E}_{\beta }(\lambda =0.144)+0.02$, $\mathcal {E}_{\beta }(\lambda =0.4)+0.05$, $\mathcal {E}_{\beta }(\lambda =0.784)+0.1$, $\mathcal {E}_{\beta }(\lambda = 1.296)+0.15$. The legend in (a) applies to all four panels.

Figure 3

Figure 4. Snapshots of QG mean vorticity fields, $\nabla ^2\psi$, at statistically steady states with (a) $\lambda =0.144$, (b) $\lambda =0.4$, (c) $\lambda =0.784$ and (d) $\lambda =1.296$. The mean energy injection rate is $\epsilon _{QG}=8\times 10^{-5}$.

Figure 4

Figure 5. The QG mean energy spectrum without NIWs at a statistically steady state with $\epsilon _{QG}=8\times 10^{-5}$.

Figure 5

Figure 6. The NIW-modified QG mean and NIW energy spectra at statistically steady states with $\lambda$ ranging from $0$ to $1.296$: (a) $E_Z(K)$, (b) $E_R(K)$, (c) $E_T(K)$ and (d) $E_W(K)$. The legend in (c) applies to all four panels. The black dash-dotted lines are power functions for reference. The black arrows show the direction of increasing $\lambda$.

Figure 6

Figure 7. The QG mean energy flux without NIWs at a statistically steady state with $k_{f}=32$ and $\epsilon _{QG}=8\times 10^{-5}$. The negative and positive values denote the upscale and downscale energy fluxes, respectively.

Figure 7

Figure 8. Energy flux of QG mean flow in NIW-containing cases of (a) $\lambda =0.144$, (b) $\lambda =0.4$, (c) $\lambda =0.784$ and (c) $\lambda =1.296$ with $k_{f}=32$ and $\epsilon _{QG}=8\times 10^{-5}$.

Figure 8

Figure 9. Ratio between jet energy and the QG mean-flow energy with different mean energy injection rates. In (b), the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

Figure 9

Figure 10. Ratio between the upscale energy flux and the total mean energy injection rate with different mean energy injection rates. In (b), the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

Figure 10

Figure 11. (a) Jet spacings and different definitions of Rhines scale with $k_{f}=32$, $\epsilon _{QG}=8\times 10^{-5}$ and varying $\lambda$. The tilde denotes the scale normalized by NIW-free jet spacing ($\lambda =0$). (b) Comparison between the numerically obtained jet spacing $\tilde {L}_{jet}$ and the mean-advection-related Rhines scale $\tilde {L}_{adv}$ with different $\lambda$. In these simulations $\epsilon _{QG}=8\times 10^{-5}$.

Figure 11

Figure 12. Zonal-mean profiles of (a) zonal velocity, (b) vorticity and (c) wave action with $\epsilon _{QG}=8\times 10^{-5}, \lambda =0.4$ at $t=2000$. The blue and red rectangles denote the regions with positive and negative vorticity, respectively.

Figure 12

Figure 13. Jet velocity profiles with $\epsilon _{QG}=8\times 10^{-5}$ and changing $\lambda$. Here, $\tilde {u}$ denotes the moved profiles sharing the same peak location for comparison; $\tilde {y}$ denotes the moved meridional coordinate.

Figure 13

Figure 14. The skewness of zonal-mean vorticity with different mean energy injection rates. In (b) the horizontal coordinate is normalized as $\tilde {\lambda }=\lambda (\epsilon _{QG}/\epsilon _{QG}^0)^{1/2}$.

Figure 14

Figure 15. Hovmöller diagram of the zonal-mean zonal velocity with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

Figure 15

Figure 16. Dependence of jet migration speed $\mathcal {V}$ on $\lambda$ with $\epsilon _{QG}=8\times 10^{-5}$.

Figure 16

Figure 17. Profiles of the normalized zonal-mean velocity $\bar {u}/\bar {u}_{max}$ and the normalized momentum flux convergence $\mathcal {T}/\mathcal {T}_{max}$ with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

Figure 17

Figure 18. (a) The dependences of the net momentum flux $\mathcal {I}$ on $\lambda$. (b) Numerically obtained jet migration speed $\mathcal {V}$ against the migration speed $\mathcal {V}^*$ calculated from momentum balance (6.5). Here, $\epsilon _{QG}=8\times 10^{-5}$.

Figure 18

Figure 19. Hovmöller diagram of the zonal-mean (a) vorticity and (b) NIW amplitude with $\epsilon _{QG}=8\times 10^{-5}$ and $\lambda =0.4$.

Figure 19

Figure 20. Illustration of the jet–NIW interaction.

Figure 20

Figure 21. The Lagrangian mean velocity profile $\bar {u}$ and the Stokes drift $\bar {u}^S$ with $\epsilon _{QG}=8\times 10^{-5}$ and $R=2.0$.