1. Introduction
Rossby waves are ubiquitous features of the atmosphere long known to produce oscillations of the westerly flow in the upper troposphere (e.g., Rossby 1939, 1940; Haurwitz 1940). Rossby waves owe their existence to gradients in potential vorticity (Hoskins et al. 1985). A basic review of Rossby wave dynamics can be found in Rhines (2002). Rossby wave packets are associated with extreme temperature, winds, and rainfall (e.g., Wirth et al. 2018), with the threat of these impacts potentially increasing with climate change (e.g., Kornhuber et al. 2020; Mann et al. 2018). The impact of Rossby waves on extreme precipitation is in part due to the poleward transport of water vapor (e.g., Grazzini et al. 2021), which can also generate ascent (e.g., Saffin et al. 2021).
Although synoptic-scale ascent in Rossby waves is typically weak, even a vertical velocity of 2 cm s−1, if sustained for 10 h, would lift an air parcel over 700 m, a substantial fraction of the depth of a typical dry-convective atmospheric boundary layer (ABL) in the warm season. Such gentle but persistent ascent could play a role in initiating or maintaining convective systems. However, one key issue is “How would this ascent vary with the diurnal cycle of winds and stability in the atmospheric boundary layer?” Addressing this question may provide insight into why there is a nocturnal maximum in warm season precipitation over the central United States (Kincer 1916; Means 1952; Wallace 1975; Easterling and Robinson 1985; Dai et al. 1999; Carbone and Tuttle 2008). Although nocturnal convection in the central United States is often associated with the eastward propagation of an envelope of dissipating and regenerating mesoscale convective systems that originated over higher terrain to the west during the previous afternoon (Dai et al. 1999; Carbone and Tuttle 2008), it can also develop without an obvious connection to pre-existing convection, and may be associated with weak but long-lived ascent (Weckwerth and Parsons 2006; Wilson and Roberts 2006; Reif and Bluestein 2017; Trier et al. 2017; Shapiro et al. 2018; Gebauer et al. 2018). These studies have focused on convective systems during the warm season, often in the absence of strong surface cold fronts, but in flow patterns that can include Rossby waves.
In this study, we introduce an analytical model to explore the timing, strength, and patterns of ascent/subsidence and nocturnal low-level jets (NLLJ) in a diurnally varying frictional (viscous) boundary layer when the flow aloft is a zonally propagating Rossby wave. As the case where the environment is stably stratified (with differential vertical motion generating horizontal temperature gradients) appears to be analytically intractable, we restrict attention to neutrally stratified flows without horizontal temperature contrasts. We thus focus on the first part (diurnal wind cycle) of the issue noted above. The flow in our model is temporally and spatially periodic with the period and wavelength of the Rossby wave. Diurnal variations in turbulence associated with the morning and evening boundary layer transitions (Stull 1988) are simply (crudely) accounted for by explicitly specifying a diurnally periodic height-independent eddy viscosity. Our model thus combines the main aspects of the classical Ekman (1905) and Åkerblom (1908) theories for frictional oceanic and atmospheric boundary layers, the Blackadar (1957) theory for NLLJs arising from inertial oscillations triggered by the cessation of dry-convective turbulence at sunset, and the Rossby (1939) theory for the motion of meanders of the jet stream. The model provides a convenient framework to explore flow dependencies on Rossby wave characteristics, latitude, time of sunset, and day and nighttime levels of turbulence intensity. However, because of its simplistic treatment of boundary layer processes and lack of thermal stratification and baroclinicity, the model should not be used for more than a qualitative view of the flow in real ABLs.
For the simpler case where the flow is horizontally uniform and the pressure gradient force is temporally constant (no Rossby wave), Blackadar (1957) proposed that the shutdown of turbulence at sunset would trigger an inertial oscillation (IO) of the ageostrophic wind, with a wind speed maximum attained when the ageostrophic wind (which rotates anticyclonically around the geostrophic wind on a hodograph diagram) aligns with the geostrophic wind. In the Blackadar theory, a strong NLLJ develops from initial (sunset) wind and geostrophic wind profiles for which the ageostrophic wind is strong. This typically occurs near Earth’s surface, where the winds are reduced from their free-atmosphere values which are often nearly geostrophic. A major role in the development of some NLLJs has been attributed to this mechanism (Parish et al. 1988; Zhong et al. 1996; Banta et al. 2002; Baas et al. 2009; Kallistratova and Kouznetsov 2012; Parish 2016, 2017; Parish and Clark 2017; Parish et al. 2020). Horizontal convergence of NLLJ winds likely contributes to the above-noted nocturnal maximum in warm season precipitation over the central United States (Pitchford and London 1962; Bonner 1966; Bonner et al. 1968; Maddox 1983; Trier and Parsons 1993; Walters and Winkler 2001; Tuttle and Davis 2006; Reif and Bluestein 2017; Shapiro et al. 2018; Gebauer et al. 2018; Weckwerth et al. 2019). Although NLLJ-associated convergence in this region is often found along the (usually northern) terminus of the NLLJ or where the NLLJ intersects a boundary layer convergence line, convergence can also arise on the lateral flanks of NLLJs from mechanisms that are less well understood.
In a study of a viscous version of the Blackadar model, Buajitti and Blackadar (1957) attributed the greater amplitude of their model wind speed oscillation at 30°N (versus 55°N) to the fact that the inertial frequency and diurnal frequency were equal at 30°N, but noted that, in their numerical experiments, “the tendency for resonance is actually not excessive because a large amount of viscous damping is characteristic of these layers” (p. 498) The theoretical studies of Paegle and Rasch (1973), Tan and Farahani (1998), Shibuya et al. (2014), Ingel’ (2015), and Momen and Bou-Zeid (2017) also suggested that IOs in the ABL could exhibit resonant amplification at critical latitudes of 30°N or 30°S (or near such latitudes if the inertial frequency is based on absolute vorticity instead of planetary vorticity). However, the extent to which resonance impacts real atmospheric flows is relatively unexplored and unclear. Walters et al. (2008) point out disagreements between climatological analyses on the location of the southerly jet frequency maximum over the Great Plains, with the 2-yr study of Bonner (1968) placing the maximum near the Kansas–Oklahoma border, some ∼7° north of 30°N, and the more extensive (40 years) study of Walters et al. (2008) putting the maximum in southern Texas, much closer to 30°N. However, the Walters et al. (2008) study also showed that, for subclasses of strong jets, the latitudes of peak jet frequency were closer to those indicated in the Bonner (1968) study.
The issue of resonance has been explored in more detail in the oceanographic literature. It has been suggested that upper-ocean Ekman layers in coastal waters forced by diurnally varying wind stresses associated with the land- and sea-breeze cycle, may display resonant responses near 30°N or 30°S (e.g., Shaffer 1972; Craig 1989; Simpson et al. 2002; Stockwell et al. 2004; Zhang et al. 2009, 2010; Hyder et al. 2011; Kim and Crawford 2014; Mihanović et al. 2016; Ashkenazy 2017; Vincze et al. 2019; Fearon et al. 2020). The most compelling evidence for diurnal-inertial resonance in a geophysical setting comes from some of these wind-driven Ekman layer studies. For example, working with the MITgcm general circulation model under periodic wind stress forcing, Ashkenazy (2017) found that the mean kinetic energy, upper-ocean current speed, and mixing layer depth were much larger in the simulation with the Coriolis frequency evaluated at 30°N than in simulations with Coriolis frequencies corresponding to other latitudes. Using the ROMS3 ocean model, Zhang et al. (2010) found that sea-breeze-driven upper-ocean currents in the Gulf of Mexico were much stronger and vertically extensive at the latitude reported closest to 30°N (29.7°N, northern Gulf) than at latitudes farther south in the Gulf. In an observational study using surface wind and ocean current data off the west coast of the United States, Kim and Crawford (2014) found that diurnal wind-current responses near 30°N were an order of magnitude larger than at other latitudes. Additionally, in laboratory experiments with a large rotating water tank configured with an oscillating horizontal plate used to simulate a periodic wind stress, Vincze et al. (2019) found that the Ekman depth and current speed were much larger when the oscillating plate frequency matched the Coriolis frequency.
Shapiro et al. (2018) proposed that the shutdown of turbulence at sunset, the trigger for a nocturnal IO/NLLJ in the Blackadar theory, can also trigger a weak but long-lived surge of convergent flow and ascent in the presence of a broad surface-based warm tongue. The governing equations in Shapiro et al. (2018) were the linearized Boussinesq equations of motion, thermal energy, and mass conservation for an inviscid stably stratified fluid on an f plane. The flow was described over one night as the solution to an initial value problem in which the initial (sunset) state was horizontally nondivergent and satisfied a zero-order jump model of a convective boundary layer. The resulting postsunset surge had aspects of an inertia–gravity wave response, with the (downward) propagation of phase lines indicating energy transfer away from the surface, and the zone of peak ascent gradually descending over the center of the warm tongue. Shapiro et al. (2018) also examined a simpler problem in which there was no warm tongue, but the free-atmosphere wind varied zonally as a stationary wave. Although the ascent in this second problem was much weaker than in the first problem, some of the weakness was likely due to the choice of parameter values. Specifically, with the southerly free-atmosphere geostrophic wind set to a small value (5 m s−1) and the initial southerly boundary layer wind set to a large fraction (80%) of that value, the initial ageostrophic wind was weak throughout the boundary layer, and would not be expected to generate a strong Blackadar-like IO/NLLJ or a strong initial surge. Stronger ageostrophic winds would presumably be attained with late afternoon winds that decreased on approach to the ground, as in the real atmosphere (or in a viscous model). Our present study is more closely related to this second problem, but differs from it in several important aspects. While the second model in Shapiro et al. (2018) was inviscid, limited to nighttime hours, stably stratified, and had a free-atmosphere wind in the form of a stationary geostrophically balanced wave, the present model is viscous (diurnally varying eddy viscosity), extends over multiple days and nights, is neutrally stratified (a major limitation), and has a free atmosphere wind that propagates as a Rossby wave.
The paper is arranged as follows. In section 2, we introduce the governing equations for the steady-periodic Ekman-Rossby problem (viscous problem) in which the eddy viscosity is diurnally periodic but otherwise temporally unrestricted. The simpler inviscid problem (limited to the nighttime) is solved in section 3. The viscous problem is solved in section 4. The derivations in sections 3 and 4 are fairly technical, but readers not interested in the details can safely skip them. Model resonance is explored in section 5. In section 6, the viscous solution is adapted to the case where the eddy viscosity varies as a piecewise constant function of time, with an abrupt increase at sunrise and an abrupt decrease at sunset. Examples of this latter flow type are presented in section 7. A summary and conclusions follow in section 8.
2. Formulating the steady-periodic Ekman-Rossby problem
A Rossby wave is propagating zonally on a uniform westerly current of strength U (>0) over the (flat) ground. The atmosphere is neutrally stratified and there are no horizontal temperature contrasts. Attention is restricted to the linearized steady-periodic state of laterally periodic flows on a beta plane. The flows are described in a Cartesian coordinate system in which x points east, y points north, and z is height above the ground. The corresponding velocity components are denoted by u = u(x, z, t), υ = υ(x, z, t), and w = w(x, z, t) (t is time), respectively. Although these components are treated as two-dimensional (x, z), there will be a slight parametric dependence on y through the latitude in the Coriolis parameter. The remote (z → ∞) velocity components are denoted by u∞ ≡ limz→∞u, υ∞ ≡ limz→∞υ, and w∞ ≡ limz→∞w. A complete list of symbols is provided in Table A1.
a. Governing equations
For simplicity, K is taken to be independent of height. Although observations in Ekman layers show that an assumed height independence for K is an oversimplification, there appear to be a variety of case-dependent behaviors (e.g., K increasing or decreasing with height, displaying multiple extrema), with no general agreement on what the height dependence actually is (O’Brien 1970; Agee et al. 1973; Jeričević and Večenaj 2009; Constantin and Johnson 2019). However, Constantin and Johnson (2019) found that large variations in K with height do not necessarily translate into large differences in the boundary layer winds: “Most importantly, we have shown, for any eddy viscosity that is bounded and tends to a constant finite value at high altitude, that the decay and spiralling of the flow upward is an enduring property… The overall picture of the classical Ekman spiral is unaltered by the details of the varying eddy viscosity, which is slightly surprising. We might have expected that viscosity profiles with a number of local maxima and minima, for example, would produce a significant distortion of the familiar structure of the flow; this is not the case” (p. 412). This conclusion may explain why geophysical models sometime produce qualitatively reasonable results using an assumed constant K. For example, Maas and van Haren (1987) found that a constant-K model led to a good description of North Sea current profiles at the dominant tidal frequencies; a decay of current speed with depth via a constant-K Ekman spiral has been observed in upper-ocean currents (e.g., Hunkins 1966; Weller 1981; Price et al. 1987; Chereskin 1995; Price and Sundermeyer 1999); atmospheric Ekman flows consistent with a constant K were observed in polar regions during the summer by Grachev et al. (2005) and Rysman et al. (2016); constant-K models have provided a good description of slope flows at night and a reasonable approximation of slope flows during the day (e.g., Defant 1949; Tyson 1968; Papadopoulos et al. 1997; Oerlemans 1998); Shapiro et al. (2022) found good agreement between the winds in a constant-K model of a baroclinic NLLJ over the Great Plains and lidar data.
The Rayleigh damping terms −R(u − u∞) and −R(υ − υ∞) in (1) and (2) are similar to terms included in slab and one-dimensional models of wind-driven Ekman layers in the upper ocean (e.g., Pollard and Millard 1970; Kundu 1976; Weller 1982; D’Asaro 1985; D’Asaro et al. 1995; Watanabe and Hibiya 2002; McWilliams and Huckle 2006; McWilliams et al. 2009; Park et al. 2009; Whitt and Thomas 2015; Gough et al. 2016; Jing et al. 2017). Weller (1982) and McWilliams et al. (2009) described the Rayleigh damping constant R as a parameterization of the momentum damping rate for upper-ocean currents due to the radiation of inertia–gravity waves to the deep ocean. As inertia–gravity waves could not be explicitly simulated in those models, the momentum damping rate associated with them could only be parameterized; inclusion of Rayleigh damping terms with time scales R−1 estimated from observations led to much improved predictions of current speeds. As inertia–gravity waves are ubiquitous in the atmosphere and can originate from the ABL (Shibuya et al. 2014; Jia et al. 2019) but cannot be simulated by our neutrally stratified model, we have included the damping terms. As we will see, Rayleigh damping precludes the development of resonant singularities in our model.
b. Boundary and temporal periodicity conditions
c. Complex wind deviation
3. Sidelight: Inviscid nocturnal flow calculation
Before solving the viscous problem of section 2, we consider the simpler problem where K vanishes at sunset, and remains zero through the night, as in the Blackadar (1957) model (although, unlike that model, our pressure gradient varies with x and t). The inviscid solutions for u, υ, and the horizontal divergence of the wind field (divergence) are simple enough that they lead to explicit formulas for the locations and magnitudes of the extrema of those variables. The inviscid flow will be compared with the viscous flow during the night in section 7.
4. Viscous solution
The solution procedure can be summarized as:
5. Solution breakdown/resonance and critical latitudes
Critical latitudes ϕc vs wavelength in the Northern Hemisphere predicted from (83). There are three critical latitudes for u and υ (corresponding to j = −1, 0, and 1) but only two critical latitudes for w (corresponding to j = −1 and j = 1).
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
6. Solution for K(t) with piecewise constant diurnal variations
7. Examples
We present examples of the piecewise-constant K(t) flows considered in section 6. There are two main experiments: (i) L2000, for an eastward-propagating wave of relatively short wavelength (L = 2000 km), and (ii) L6000, for a westward-propagating wave of moderate wavelength (L = 6000 km). In both experiments, we take ϕ = 40°N (
Input parameters for experiments L2000 and L6000. These are also the default values for further experiments in which one parameter is varied at a time.
Westerly current speed U vs the latitude ϕ obtained from (13) using the parameters in Table 1. UL2000 and UL6000 are the values of U from experiments in which L = 2000 km and L = 6000 km, respectively.
Horizontal contour plots of Π (m2 s−2) at t = 0 in experiments (a) L2000 and (b) L6000. In view of (21), isolines of Π drawn in the x–y plane coincide with remote (z → ∞) flow streamlines.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Hovmöller plots of Π, u, υ, and w from experiment L2000 are shown in Fig. 3; u and υ are shown at the heights of their maxima (zumax ≈ 375 m, zυmax ≈ 475 m), w is shown at z = 1000 m, which is within or near the top of typical ABLs, and the Π panel is valid at all heights (recall that Π is independent of z). All fields except Π exhibit marked diurnal variations. There is a postsunset surge of alternating (with respect to x) convergent and divergent flow. The magnitudes and gradients of u, υ, and w are weaker during the day than during the night (although this feature does not hold for w at higher levels; the height dependence of w will be discussed later). Consistent with the presence of a westerly current, the positive (westerly) peak value of u is larger than the magnitude of its largest negative (easterly) value of u. The peak westerly and easterly winds are found in the regions of strongest westerly- and easterly-directed pressure gradient forces, respectively. From the locations of the u-extrema we infer that the peak horizontal convergence (−∂u/∂x) is largest along the trough, where w attains its maximum. From the locations of the υ-extrema we infer that the peak vertical vorticity (∂υ/∂x) is also found along the trough. This finding is consistent with the formula for Ekman pumping [e.g., (4.5.39) of Pedlosky (1987)] in which w is directly proportional to the vertical vorticity.
Hovmöller plots from experiment L2000: (a) Π (m2 s−2), (b) w (cm s−1) at z = 1000 m, (c) u (m s−1) at height of its maximum (zumax), and (d) υ (m s−1) at height of its maximum (zυmax); zumax ≈ 375 m, and zυmax ≈ 475 m. Thick dashed yellow and black lines indicate times of sunrise and sunset, respectively. Thick solid black curves track the locations of extrema (with respect to x) predicted by the inviscid theory for the nighttime intervals.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Figure 3 also shows that NLLJ winds begin to develop shortly after sunset and reach peak intensity a few hours before sunrise. These winds undergo an anticyclonic rotation through the night (e.g., during the first night, the winds near the center of the domain veer from southeasterly shortly after sunset to southwesterly by sunrise). While such NLLJs are consistent with the inviscid Blackadar IO mechanism, they cannot be pure IOs since the pressure gradient force is time dependent. However, the u and υ fields are in good agreement with the u and υ predicted by our inviscid theory. For example, the locations of the extrema of u, υ, and w are close to the locations xu(t), xυ(t), and xdiv(t) given in (41), (45), and (43), respectively (solid black curves in Fig. 3). Additionally, the peak nighttime values of the inviscid solutions (42) for upeak (≈23.3 m s−1) and (46) for υpeak (≈31.0 m s−1) closely match the peak values from the model (umax ≈ 22.6 m s−1, υmax ≈ 31.2 m s−1). The wind and vertical velocity fields weaken rapidly after sunrise.
The Hovmöller plots from experiment L6000 (Fig. 4) display many of the features seen in Fig. 3 for L2000, including a postsunset surge of convergent/divergent flow, the development of strong NLLJs with anticyclonically turning winds, and the rapid decay of u, υ, and w after sunrise (presumably the changes to u, υ, and w after sunrise and sunset would not have been as rapid if K had been specified to change gradually during the transition periods). The υmax in L6000 (≈30.0 m s−1) is close to its value in L2000 (≈31.2 m s−1). The somewhat larger difference between the umax in L6000 (≈19.0 m s−1) and L2000 (≈22.6 m s−1) is consistent with the U obtained from the dispersion relation at 40°N being about 3.5 m s−1 larger in L2000 than in L6000 (Table 2). Again, our inviscid theory is in good agreement with the viscous model during the nighttime hours; the tracks of the peaks in the model-predicted u, υ, and w are well represented by the inviscid curves, and the peak values from the inviscid solutions (42) and (46) (upeak ≈ 17.7 m s−1, υpeak ≈ 29.6 m s−1) are close to the above-noted values for umax and υmax.
As in Fig. 3, but for experiment L6000. Here, zumax ≈ 350 m, and zυmax ≈ 500 m. Note that color bar bin intervals are different from those in Fig. 3.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Vertical cross sections of u, υ, and w through the x locations of their maxima (though for w only the location of the nocturnal maximum is considered) are shown for L2000 in Fig. 5 and for L6000 in Fig. 6. Vertical cross sections of w through their daytime maxima (not shown) are fairly similar to these latter plots of w. The υ panels show the development of NLLJs with peak υ more than 50% greater than the peak remote value (A = 20 m s−1). The peak jet winds in these experiments are found at z = 600 m (in L2000) and z = 625 m (in L6000). These heights are consistent with values documented in climatological studies; NLLJ winds typically peak at heights less than 1 km, and frequently at levels of 500 m or less (e.g., Whiteman et al. 1997; Song et al. 2005; Baas et al. 2009; Carroll et al. 2019). The w panels show two modes of ascent, an elevated mode that peaks in mid/late-afternoon and a lower-level mode that peaks a few hours after sunset. Although the afternoon peak is stronger than the nocturnal peak, the nocturnal peak is stronger than the daytime peak for z < 2000 m (recall the postsunset surge of w seen in Figs. 3b and 4b at z = 1000 m).
Time–height plots of (a) u (m s−1), (b) υ (m s−1), and (c) w (cm s−1) at the x locations of their respective domain-wide maxima (though only the x location of nighttime maximum is considered for w) in experiment L2000. Thick dashed yellow and black lines indicate times of sunrise and sunset, respectively.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
As in Fig. 5, but for experiment L6000. Note that color bar bin intervals are different from those in Fig. 5.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Time series of the domain-maximum values umax, υmax, and wmax from L2000 and L6000 are shown in Figs. 7 and 8 respectively.3 Although none of u, υ, or w is diurnally periodic (see Figs. 3–6), Figs. 7 and 8 show that umax, υmax, and wmax are diurnally periodic. Not surprisingly, the peak remote values of u (i.e., the U given in Table 2 for 40°N: U ≈ 9.5 m s−1 in L2000 and U ≈ 6.0 m s−1 in L6000) and of υ (A = 20 m s−1) serve as floors for umax and υmax over the wave period. The umax and υmax panels show the NLLJ developing after sunset and decaying shortly after sunrise. The wmax panels show the main double-mode of ascent along with a weak third peak that arises when the late-night trend for w to increase is reversed by the sudden weakening of the convergent flow at sunrise. The overall peak wmax for L = 2000 km (≈5.3 cm s−1) is ≈2.5 times greater than that for L = 6000 km (≈2.1 cm s−1). This dependence of ascent on wavelength is slightly weaker than the factor of 3 increase implied by (76). The discrepancy may be due to the fact that the condition for the validity of (76) (kU/f ≪ 1) is only marginally satisfied for L = 2000 km (kU/f is ≈0.067 for L = 6000 km but ≈0.318 for L = 2000 km).
Time series of u, υ, and w maxima from experiment L2000: (a) umax (m s−1), (b) υmax (m s−1), and (c) wmax (cm s−1). Also shown are results from L2000-like experiments with no diurnal variations (K is constant); solid blue lines for K = 2 m2 s−1, and dashed orange lines for K = 50 m2 s−1. The maxima are the largest values in the analysis domain (one wavelength, from ground to z = 10 km).
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
As in Fig. 7, but for experiment L6000. Tick mark intervals/labels differ from those in Fig. 7.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Figures 7 and 8 also display time series from experiments in which K has no diurnal variations. We consider K = 50 m2 s−1 (former daytime-only value) in one experiment, and K = 2 m2 s−1 (former nighttime-only value) in another. Although the u and υ in these constant-K experiments are temporally periodic with frequency ω (figures not shown), Figs. 7 and 8 show that umax and υmax are temporally constant. Notably, for both L2000- and L6000-like experiments, the umax and υmax obtained with K = 50 m2 s−1 are the same as those obtained with K = 2 m2 s−1. The K-independence of these maxima is a feature of the classical Ekman solution [can be inferred from (4.3.21a) and (4.3.21b) of Pedlosky (1987)]. Additionally, consistent with the Ekman prediction that the heights of the wind maxima are proportional to the Ekman depth (2K/f)1/2, the ratio of the heights of the peak u for the two constant-K L6000-like runs (1125 m/225 m = 5.0) is nearly the same as that ratio in the L2000-like runs (1275 m/250 m = 5.1) and the ratio of the square roots of the two K (
The dependence of umax, υmax, and wmax on latitude is shown in Fig. 9 for L2000- and L6000-like experiments. While there is relatively little change in umax or υmax with latitude, wmax does peak at the critical latitudes predicted in (83) (see also Fig. 1) for j = −1 and j = 1, particularly for L = 2000 km. Figure 9 also presents results from runs in which the Rayleigh damping parameter is reduced to (30 day)−1. For both wavelengths, the umax and υmax in the 30-day damping runs are nearly the same as in the 5-day damping runs for all latitudes. However, while wmax in the 30- and 5-day damping runs are similar for most latitudes, in the 30-day damping runs there is a marked increase in wmax on approach to the j = −1 and j = 1 critical latitudes, particularly for the L = 2000 km runs. Indeed, wmax at the critical latitude of 32.8°N is ≈8.0 cm s−1 with 5-day damping but ≈12.0 cm s−1 with 30-day damping.
The u, υ, and w maxima as functions of latitude in experiments L2000 and L6000 and in related experiments in which the damping parameter R provides 30-day damping instead of 5-day damping: (a) umax (m s−1), (b) υmax (m s−1), and (c) wmax (cm s−1). Results are shown for (blue line) experiment L2000, (orange line) an L2000-like run with 30-day damping, (green line) experiment L6000, and (red line) an L6000-like run with 30-day damping. The maxima are the largest values in the analysis domain over one wave period. Dashed vertical lines in (c) mark the critical latitudes predicted by (83) (see also Fig. 1).
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Wind deviation depth scale h as a function of latitude in experiments (a) L2000 and (b) L6000. Also shown are results from experiments in which the damping parameter was adjusted to provide 30-day damping instead of 5-day damping: Blue lines depict results from experiments L2000 and L6000. Orange lines depict results from L2000- and L6000-like runs with 30-day damping. Dashed vertical lines mark the critical latitudes predicted by (83) (see also Fig. 1).
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Time series of u, υ, and w in L2000-like experiments at the 32.82°N critical latitude and a latitude just ∼3° away (36°N) are shown on the z = 5 km surface for cases of 5- and 30-day damping in Fig. 11. The differences between the velocity components in the 5- and 30-day damping runs are much larger at 32.82°N than at 36°N. The amplitude of the w-oscillation and the peak value of w in the 30-day damping runs is much larger at 32.82°N; from t = 34 h to t = 40 h and from t = 54 h to t = 62 h, w is over 3 cm s−1 larger (40%–50% greater) at 32.82°N than at 36°N. Discrepancies between the υ fields at the two latitudes exceed 5 m s−1 but appear to be more associated with a phase shift than an amplification.
Time series of (a) u (m s−1), (b) υ (m s−1), and (c) w (cm s−1) at z = 5 km at the x locations of their respective domain-wide maxima (though only the x location of nighttime maximum is considered for w). Results are shown for L2000-like runs at the 32.82°N critical latitude (blue lines) and at 36°N (red lines) for 5-day damping (solid lines) and 30-day damping (dashed lines).
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Time–height plots of w at different latitudes for both L2000- and L6000-like runs show that the daytime peak (local maximum) of w occurs increasingly later in the afternoon at successively lower latitudes until, at sufficiently small latitudes, the afternoon peak disappears. The nocturnal mode dominates south of the 32.8°N critical latitude for L = 2000 km and south of the 38.2°N critical latitude for L = 6000 km. This behavior can be seen in Fig. 12 for L2000-like runs at 50° and 30°N (cf. with Fig. 5c for 40°N).
Time–height plots of w (cm s−1) through the x locations of their nocturnal maxima for L2000-like runs at (a) 50° and (b) 30°N; see Fig. 5c for the 40°N plot. Thick dashed yellow and black lines indicate times of sunrise and sunset, respectively. Note the differences between color bars.
Citation: Journal of the Atmospheric Sciences 81, 3; 10.1175/JAS-D-23-0070.1
Last, L2000- and L6000-like experiments were run to examine the sensitivity of the flow to the time of sunset at different latitudes. In the L2000-like runs, the changes in both umax and υmax between an early sunset of tset = 11 h and a late sunset of tset = 15 h are less than 0.7 m s−1 (relative changes < 3%) throughout a 30° to 50°N latitude band. The corresponding relative changes in wmax, although larger than the relative changes in umax and υmax, are still small. For example, at 30°N, wmax decreases from ≈7.9 cm s−1 when tset = 11 h to ≈6.5 cm s−1 when tset = 15 h (18% decrease), while at 50°N, the corresponding change in wmax is an increase from ≈3.9 to ≈4.6 cm s−1 (18% increase). Qualitatively similar relative changes in the wind and vertical velocity maxima are found in the L6000-like runs. Unsurprisingly, for both wavelengths, the winds and vertical velocity attain their nighttime maxima at roughly the same time relative to sunset and thus occur later with later sunsets. In cases where the overall peak vertical velocity occurs during the daytime, the time of the maximum also occurs later with later sunsets, but the shift is not as pronounced as for the nocturnal maximum (e.g., there is only a 1.4 h shift in the afternoon maximum in w from early to late sunsets at 30°N with L = 6000 km).
8. Summary and conclusions
The linkage between Rossby waves and extreme weather events, including heavy rainfall and flooding, is well established (Wirth et al. 2018). As convection is often impacted by conditions in the lower troposphere, we are motivated to study variations of wind and vertical velocity in the boundary layer during the passage of a zonally propagating Rossby wave. An analytical model is introduced to explore the spatial and temporal characteristics of NLLJs and ascent (Ekman pumping) in a diurnally varying frictional boundary layer during wave passage. The model combines the main aspects of the Ekman (1905), Åkerblom (1908) theories for frictional boundary layers on the rotating Earth, the Blackadar (1957) theory for NLLJs arising from inertial oscillations triggered by the shutdown of turbulence at sunset, and the Rossby (1939) theory for barotropic Rossby waves on a beta plane. The governing equations are the linearized Reynolds-averaged Boussinesq- and beta-plane-approximated horizontal equations of motion, the hydrostatic equation, and the incompressibility condition. The restrictions to a barotropic setting and a neutrally stratified environment are major limitations of our model. As in the slab and one-dimensional models of wind-induced Ekman layers in the upper ocean discussed in the introduction, our model includes Rayleigh damping terms to account for momentum damping associated with inertia–gravity waves that cannot be explicitly simulated. Since our model is restricted to the steady-periodic state, the Rossby wave period must be an integral number of days (diurnal frequency is a harmonic of the wave frequency). The governing equations are solved for eddy viscosities K that vary diurnally but are independent of height. We consider diurnally varying K with otherwise arbitrary time variations and K that vary as piecewise constant functions of time with step changes at sunrise and sunset. Flows of the latter type are explored for relatively short (L = 2000 km) and moderate (L = 6000 km) wavelengths. Additionally, these viscous flows are compared (at night) to flows obtained from an inviscid calculation. Among the model predictions are the following:
-
NLLJ winds begin to develop at sunset, turn anticyclonically through the night, and reach peak strength a few hours after midnight. Such flows are consistent with the inviscid Blackadar (1957) IO mechanism for NLLJs, but are not pure IOs since our pressure gradient force varies in x and t. The magnitudes and locations of the peak u, υ, and horizontal divergence fields from the viscous model are in good agreement with those from the inviscid calculation during the nighttime. The extrema in u and in the horizontal divergence of the wind field move with the westerly current.
-
There are two main modes of ascent: an elevated daytime mode and a lower-altitude nighttime mode. The nighttime mode is associated with a weak but persistent surge of convergent flow triggered by the shutdown of turbulence at sunset, the same mechanism that triggers IOs/NLLJs in the Blackadar model. The nocturnal surge is reminiscent of that predicted in Shapiro et al. (2018), although the inertia–gravity wave characteristics of the ascent in that study cannot be simulated in the present neutrally stratified model. The daytime peak w occurs increasingly later in the afternoon with decreasing latitude. South of critical latitudes (32.8°N for L = 2000 km and 38.2°N for L = 6000 km), the nighttime mode dominates. The model predicts that w ∝ kA for Rossby numbers kU/f ≪ 1.
-
If the Rayleigh damping is turned off (R = 0), the model breaks down where (77) is satisfied, that is, where the inertial frequency, advection frequency (jkU; j = −1, 0, 1), and a harmonic of the wave frequency (of which the diurnal frequency is an integral multiple) sum to 0. This occurs at the three critical latitudes associated with j = −1, 0, 1 in (83) for the Northern Hemisphere. Corresponding to j = 0 is the 30°N diurnal-inertial resonant latitude noted in previous studies (see the introduction). The critical latitudes associated with j = −1 and j = 1 vary with wavelength. Breakdown is manifested as (i) u and υ fields that do not approach their free-atmosphere profiles as z → ∞ at the three critical latitudes, in violation of a remote boundary condition, and (ii) a vertical velocity that becomes infinite as z → ∞ at the j = −1 and j = 1 critical latitudes. However, there is no prediction for u or υ to peak at any critical latitude.
-
Breakdown cannot occur if R ≠ 0. However, with both R = (5 days)−1 and R = (30 days)−1 damping, the peak w is found at the j = −1 and j = 1 critical latitudes. The effect is most pronounced for the run with shorter wavelength (L = 2000 km) and weaker damping [R = (30 days)−1], where w increases from ≈4.5 cm s−1 at 50°N to ≈8.0 cm s−1 at the 32.8°N critical latitude. Additionally, a wind deviation depth scale h [defined via (98) and (99)] is significantly larger at the three critical latitudes for both L = 2000 km and L = 6000 km when the damping is weak [R = (30 days)−1]; the response is muted for the more strongly damped case [R = (5 days)−1]. The tendency for the depth scale to increase at the critical latitudes in the weakly damped case was reminiscent of the increased upper-ocean mixed layer depths at the 30°N critical latitude for diurnal-inertial resonance documented in the oceanographic literature (e.g., Zhang et al. 2010; Ashkenazy 2017; Vincze et al. 2019). In contrast to those studies, however, no discernible increases in our model u or υ fields were found in the vicinity of any critical latitude; for L = 2000 km, u and υ are remarkably uniform across a 20°–50°N latitude band while, for L = 6000 km, υ is again nearly uniform across that band but u decreases northward, a trend consistent with the northward decrease in U required by the dispersion relation (see Table 2).
The vertical velocities predicted by our diurnally varying Ekman-Rossby model, particularly for the waves of smaller wavelength (L = 2000 km), can plausibly play a role in the initiation or maintenance of deep convection. Our study was motivated by questions about the nocturnal convection maximum observed during the warm season over the Great Plains. In this context, the ascent after sunset may facilitate the intensification or persistence of eastward-propagating convection that initiated during the afternoon over higher terrain to the west. However, without provision for an ambient stratification, the model’s vertical velocities are likely overestimates (no lid effect) and it is not possible to study any direct role of inertia–gravity waves in countering a resonant response. As our analytical model framework cannot be extended to include stratification (via a thermal energy equation) since separable solutions are not possible for that case, numerical model simulations may be the best tool to explore the more realistic stratified problem. Numerical simulations can also be used to explore possible synergies between the ascent mechanism described in this paper and ascent associated with warm tongues or other surface-based thermal inhomogeneities typically found over the Great Plains, and the extent to which the combined effects increase the likelihood for deep convection.
Although the Rossby wave in our study has a single frequency, that frequency can be thought of, more broadly, as representing the dominant Fourier component in a continuous power density spectrum.
As preliminary experiments using Nτ = 7 but sgnω = 1 (westerly propagating wave) yielded values of U that were unrealistically large, we have restricted attention to the retrogressing case.
The same symbol for the maximum value of a variable (e.g. umax) is used to represent both the maximum value at a given time and the maximum value over the wave period. It should be clear from the context which type of maximum is being considered.
The analogous υ-based depth scale h differed from this u-based scale by <200 m for L = 2000 km and by <450 m for L = 6000 km for latitudes between 15° and 50°N.
Acknowledgments.
This research was supported by NSF AGS-1921587. This paper has benefited greatly from the many helpful comments of the anonymous referees. We also thank Hristo Chipilski for his comments on an earlier presentation on this topic. Shawn Riley and David Goines provided computer assistance.
Data availability statement.
The analytical model is available from the lead author upon request.
APPENDIX A
List of Symbols
Table A1 provides a list of symbols used in the article.
Symbols, their descriptions, and first numbered equations in which they appear (definitions usually given nearby). If a symbol does not appear in any numbered equation, the section where it is first used is noted.
APPENDIX B
Uniqueness Proof
APPENDIX C
Inviscid υ Extrema
REFERENCES
Agee, E. M., D. E. Brown, T. S. Chen, and K. E. Dowell, 1973: A height-dependent model of eddy viscosity in the planetary boundary layer. J. Appl. Meteor., 12, 409–412, https://doi.org/10.1175/1520-0450(1973)012<0409:AHDMOE>2.0.CO;2.
Åkerblom, F., 1908: Rechercher sur les courants le plus bas de l’atmosphere au-dessus de Paris. Nova Acta Regiae Soc. Sci. Ups., 2, 203–251.
Ashkenazy, Y., 2017: Energy transfer of surface wind-induced currents to the deep ocean via resonance with the Coriolis force. J. Mar. Syst., 167, 93–104, https://doi.org/10.1016/j.jmarsys.2016.11.019.
Baas, P., F. C. Bosveld, H. Klein Baltink, and A. A. M. Holtslag, 2009: A climatology of nocturnal low-level jets at Cabauw. J. Appl. Meteor. Climatol., 48, 1627–1642, https://doi.org/10.1175/2009JAMC1965.1.
Banta, R. M., R. K. Newsom, J. K. Lundquist, Y. L. Pichugina, R. L. Coulter, and L. Mahrt, 2002: Nocturnal low-level jet characteristics over Kansas during CASES-99. Bound.-Layer Meteor., 105, 221–252, https://doi.org/10.1023/A:1019992330866.
Blackadar, A. K., 1957: Boundary layer wind maxima and their significance for the growth of nocturnal inversions. Bull. Amer. Meteor. Soc., 38, 283–290, https://doi.org/10.1175/1520-0477-38.5.283.
Bonner, W. D., 1966: Case study of thunderstorm activity in relation to the low-level jet. Mon. Wea. Rev., 94, 167–178, https://doi.org/10.1175/1520-0493(1966)094<0167:CSOTAI>2.3.CO;2.
Bonner, W. D., 1968: Climatology of the low level jet. Mon. Wea. Rev., 96, 833–850, https://doi.org/10.1175/1520-0493(1968)096<0833:COTLLJ>2.0.CO;2.
Bonner, W. D., S. Esbensen, and R. Greenberg, 1968: Kinematics of the low-level jet. J. Appl. Meteor., 7, 339–347, https://doi.org/10.1175/1520-0450(1968)007<0339:KOTLLJ>2.0.CO;2.
Buajitti, K., and A. K. Blackadar, 1957: Theoretical studies of diurnal wind-structure variations in the planetary boundary layer. Quart. J. Roy. Meteor. Soc., 83, 486–500, https://doi.org/10.1002/qj.49708335804.
Carbone, R. E., and J. D. Tuttle, 2008: Rainfall occurrence in the U.S. warm season: The diurnal cycle. J. Climate, 21, 4132–4146, https://doi.org/10.1175/2008JCLI2275.1.
Carroll, B. J., B. B. Demoz, and R. Delgado, 2019: An overview of low-level jet winds and corresponding mixed layer depths during PECAN. J. Geophys. Res. Atmos., 124, 9141–9160, https://doi.org/10.1029/2019JD030658.
Chereskin, T. K., 1995: Direct evidence for an Ekman balance in the California Current. J. Geophys. Res., 100, 18 261–18 269, https://doi.org/10.1029/95JC02182.
Constantin, A., and R. S. Johnson, 2019: Atmospheric Ekman flows with variable eddy viscosity. Bound.-Layer Meteor., 170, 395–414, https://doi.org/10.1007/s10546-018-0404-0.
Craig, P. D., 1989: A model of diurnally forced vertical current structure near 30° latitude. Cont. Shelf Res., 9, 965–980, https://doi.org/10.1016/0278-4343(89)90002-2.
Dai, A., F. Giorgi, and K. E. Trenberth, 1999: Observed and model-simulated diurnal cycles of precipitation over the contiguous United States. J. Geophys. Res., 104, 6377–6402, https://doi.org/10.1029/98JD02720.
Dandou, A., M. Tombrou, K. Schäfer, S. Emeis, A. P. Protonotariou, E. Bossioli, N. Soulakellis, and P. Suppan, 2009: A comparison between modelled and measured mixing-layer height over Munich. Bound.-Layer Meteor., 131, 425–440, https://doi.org/10.1007/s10546-009-9373-7.
D’Asaro, E. A., 1985: The energy flux from the wind to near-inertial motions in the surface mixed layer. J. Phys. Oceanogr., 15, 1043–1059, https://doi.org/10.1175/1520-0485(1985)015<1043:TEFFTW>2.0.CO;2.
D’Asaro, E. A., C. C. Eriksen, M. D. Levine, C. A. Paulson, P. Niiler, and P. V. Meurs, 1995: Upper-ocean inertial currents forced by a strong storm. Part I: Data and comparisons with linear theory. J. Phys. Oceanogr., 25, 2909–2936, https://doi.org/10.1175/1520-0485(1995)025<2909:UOICFB>2.0.CO;2.
Defant, F., 1949: Zur theorie der Hangwinde, nebst Bemerkungen zur theorie der Berg- und Talwinde. Arch. Meteor. Geophys. Bioklimatol., A1, 421–450, https://doi.org/10.1007/BF02247634.
Easterling, D. R., and P. J. Robinson, 1985: The diurnal variation of thunderstorm activity in the United States. J. Climate Appl. Meteor., 24, 1048–1058, https://doi.org/10.1175/1520-0450(1985)024<1048:TDVOTA>2.0.CO;2.
Ekman, V. W., 1905: On the influence of the Earth’s rotation on ocean-currents. Arch. Math. Astron. Phys., 2 (11), 1–52.
Fearon, G., S. Herbette, J. Veitch, G. Cambon, A. J. Lucas, F. Lemarié, and M. Vichi, 2020: Enhanced vertical mixing in coastal upwelling systems driven by diurnal-inertial resonance: Numerical experiments. J. Geophys. Res. Oceans, 125, e2020JC016208, https://doi.org/10.1029/2020JC016208.
Gebauer, J. G., A. Shapiro, E. Fedorovich, and P. Klein, 2018: Convection initiation caused by heterogeneous low-level jets over the Great Plains. Mon. Wea. Rev., 146, 2615–2637, https://doi.org/10.1175/MWR-D-18-0002.1.
Gough, M. K., A. J. H. M. Reniers, J. H. MacMahan, and S. D. Howden, 2016: Resonant near-surface inertial oscillations in the northeastern Gulf of Mexico. J. Geophys. Res. Oceans, 121, 2163–2182, https://doi.org/10.1002/2015JC011372.
Grachev, A. A., C. W. Fairall, P. O. G. Persson, E. L. Andreas, and P. S. Guest, 2005: Stable boundary-layer scaling regimes: The SHEBA data. Bound.-Layer Meteor., 116, 201–235, https://doi.org/10.1007/s10546-004-2729-0.
Grazzini, F., G. Fragkoulidis, F. Teubler, V. Wirth, and G. C. Craig, 2021: Extreme precipitation events over northern Italy. Part II: Dynamical precursors. Quart. J. Roy. Meteor. Soc., 147, 1237–1257, https://doi.org/10.1002/qj.3969.
Gustafson, K. E., 1987: Introduction to Partial Differential Equations and Hilbert Space Methods. 2nd ed. John Wiley and Sons, 409 pp.
Haurwitz, B., 1940: The motion of atmospheric disturbances. J. Mar. Res., 3, 35–50.
Hoskins, B. J., M. E. McIntyre, and A. W. Robertson, 1985: On the use and significance of isentropic potential vorticity maps. Quart. J. Roy. Meteor. Soc., 111, 877–946, https://doi.org/10.1002/qj.49711147002.
Hunkins, K., 1966: Ekman drift currents in the Arctic Ocean. Deep-Sea Res., 13, 607–620, https://doi.org/10.1016/0011-7471(66)90592-4.
Hyder, P., J. H. Simpson, J. Xing, and S. T. Gille, 2011: Observations over an annual cycle and simulations of wind-forced oscillations near the critical latitude for diurnal-inertial resonance. Cont. Shelf Res., 31, 1576–1591, https://doi.org/10.1016/j.csr.2011.06.001.
Ingel’, L. K., 2015: One type of resonance phenomena in the atmosphere and water bodies. Fluid Dyn., 50, 494–500, https://doi.org/10.1134/S0015462815040043.
Jeričević, A., and Ž. Večenaj, 2009: Improvement of vertical diffusion analytic schemes under stable atmospheric conditions. Bound.-Layer Meteor., 131, 293–307, https://doi.org/10.1007/s10546-009-9367-5.
Jia, M., and Coauthors, 2019: Long-lived high-frequency gravity waves in the atmospheric boundary layer: Observations and simulations. Atmos. Chem. Phys., 19, 15 431–15 446, https://doi.org/10.5194/acp-19-15431-2019.
Jing, Z., L. Wu, and X. Ma, 2017: Energy exchange between the mesoscale oceanic eddies and wind-forced near-inertial oscillations. J. Phys. Oceanogr., 47, 721–733, https://doi.org/10.1175/JPO-D-16-0214.1.
Kallistratova, M. A., and R. D. Kouznetsov, 2012: Low-level jets in the Moscow region in summer and winter observed with a Sodar network. Bound.-Layer Meteor., 143, 159–175, https://doi.org/10.1007/s10546-011-9639-8.
Kim, S. Y., and G. Crawford, 2014: Resonant ocean current responses driven by coastal winds near the critical latitude. Geophys. Res. Lett., 41, 5581–5587, https://doi.org/10.1002/2014GL060402.
Kincer, J. B., 1916: Daytime and nighttime precipitation and their economic significance. Mon. Wea. Rev., 44, 628–633, https://doi.org/10.1175/1520-0493(1916)44<628:DANPAT>2.0.CO;2.
Kornhuber, K., D. Coumou, E. Vogel, C. Lesk, J. F. Donges, J. Lehmann, and R. M. Horton, 2020: Amplified Rossby waves enhance risk of concurrent heatwaves in major breadbasket regions. Nat. Climate Change, 10, 48–53, https://doi.org/10.1038/s41558-019-0637-z.
Kundu, P. K., 1976: An analysis of inertial oscillations observed near Oregon coast. J. Phys. Oceanogr., 6, 879–893, https://doi.org/10.1175/1520-0485(1976)006<0879:AAOIOO>2.0.CO;2.
Maas, L. R. M., and J. J. M. van Haren, 1987: Observations on the vertical structure of tidal and inertial currents in the central North Sea. J. Mar. Res., 45, 293–318, https://doi.org/10.1357/002224087788401106.
Maddox, R. A., 1983: Large-scale meteorological conditions associated with midlatitude mesoscale convective complexes. Mon. Wea. Rev., 111, 1475–1493, https://doi.org/10.1175/1520-0493(1983)111<1475:LSMCAW>2.0.CO;2.
Mahrt, L., and D. Vickers, 2006: Extremely weak mixing in stable conditions. Bound.-Layer Meteor., 119, 19–39, https://doi.org/10.1007/s10546-005-9017-5.
Mann, M. E., S. Rahmstorf, K. Kornhuber, B. A. Steinman, S. K. Miller, S. Petri, and D. Coumou, 2018: Projected changes in persistent extreme summer weather events: The role of quasi-resonant amplification. Sci. Adv., 4, eaat3272, https://doi.org/10.1126/sciadv.aat3272.
McWilliams, J. C., and E. Huckle, 2006: Ekman layer rectification. J. Phys. Oceanogr., 36, 1646–1659, https://doi.org/10.1175/JPO2912.1.
McWilliams, J. C., E. Huckle, and A. F. Shchepetkin, 2009: Buoyancy effects in a stratified Ekman layer. J. Phys. Oceanogr., 39, 2581–2599, https://doi.org/10.1175/2009JPO4130.1.
Means, L. L., 1952: On thunderstorm forecasting in the central United States. Mon. Wea. Rev., 80, 165–189, https://doi.org/10.1175/1520-0493(1952)080<0165:OTFITC>2.0.CO;2.
Mihanović, H., C. Pattiaratchi, and F. Verspecht, 2016: Diurnal sea breezes force near-inertial waves along Rottnest continental shelf, southwestern Australia. J. Phys. Oceanogr., 46, 3487–3508, https://doi.org/10.1175/JPO-D-16-0022.1.
Momen, M., and E. Bou-Zeid, 2017: Analytical reduced models for the non-stationary diabatic atmospheric boundary layer. Bound.-Layer Meteor., 164, 383–399, https://doi.org/10.1007/s10546-017-0247-0.
O’Brien, J. J., 1970: A note on the vertical structure of the eddy exchange coefficient in the planetary boundary layer. J. Atmos. Sci., 27, 1213–1215, https://doi.org/10.1175/1520-0469(1970)027<1213:ANOTVS>2.0.CO;2.
Oerlemans, J., 1998: The atmospheric boundary layer over melting glaciers. Clear and Cloudy Boundary Layers, A. A. M. Holtslag and P. G. Duynkerke, Eds., Royal Netherlands Academy of Arts and Sciences, 129–153.
Ooyama, K., 1957: A study of diurnal variation of wind caused by periodic variation of eddy viscosity. Final Rep. AF19(604)-1368, 80–135.
Paegle, J., and G. E. Rasch, 1973: Three-dimensional characteristics of diurnally varying boundary-layer flows. Mon. Wea. Rev., 101, 746–756, https://doi.org/10.1175/1520-0493(1973)101<0746:TCODVB>2.3.CO;2.
Papadopoulos, K. H., C. G. Helmis, A. T. Soilemes, J. Kalogiros, P. G. Papageorgas, and D. N. Asimakopoulos, 1997: The structure of katabatic flows down a simple slope. Quart. J. Roy. Meteor. Soc., 123, 1581–1601, https://doi.org/10.1002/qj.49712354207.
Parish, T. R., 2016: A comparative study of the 3 June 2015 Great Plains low-level jet. Mon. Wea. Rev., 144, 2963–2979, https://doi.org/10.1175/MWR-D-16-0071.1.
Parish, T. R., 2017: On the forcing of the summertime Great Plains low-level jet. J. Atmos. Sci., 74, 3937–3953, https://doi.org/10.1175/JAS-D-17-0059.1.
Parish, T. R., and R. D. Clark, 2017: On the initiation of the 20 June 2015 Great Plains low-level jet. J. Appl. Meteor. Climatol., 56, 1883–1895, https://doi.org/10.1175/JAMC-D-16-0187.1.
Parish, T. R., A. R. Rodi, and R. D. Clark, 1988: A case study of the summertime Great Plains low level jet. Mon. Wea. Rev., 116, 94–105, https://doi.org/10.1175/1520-0493(1988)116<0094:ACSOTS>2.0.CO;2.
Parish, T. R., R. D. Clark, and T. D. Sikora, 2020: Nocturnal destabilization associated with the summertime Great Plains low-level jet. Mon. Wea. Rev., 148, 4641–4656, https://doi.org/10.1175/MWR-D-19-0394.1.
Park, J. J., K. Kim, and R. W. Schmitt, 2009: Global distribution of the decay timescale of mixed layer inertial motions observed by satellite-tracked drifters. J. Geophys. Res., 114, C11010, https://doi.org/10.1029/2008JC005216.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. 2nd ed. Springer-Verlag, 710 pp.
Pitchford, K. L., and J. London, 1962: The low-level jet as related to nocturnal thunderstorms over Midwest United States. J. Appl. Meteor., 1, 43–47, https://doi.org/10.1175/1520-0450(1962)001<0043:TLLJAR>2.0.CO;2.
Pollard, R. T., and R. C. Millard Jr., 1970: Comparison between observed and simulated wind-generated inertial oscillations. Deep-Sea Res., 17, 813–821, https://doi.org/10.1016/0011-7471(70)90043-4.
Price, J. F., and M. A. Sundermeyer, 1999: Stratified Ekman layers. J. Geophys. Res., 104, 20 467–20 494, https://doi.org/10.1029/1999JC900164.
Price, J. F., R. A. Weller, and R. R. Schudlich, 1987: Wind-driven ocean currents and Ekman transport. Science, 238, 1534–1538, https://doi.org/10.1126/science.238.4833.1534.
Reif, D. W., and H. B. Bluestein, 2017: A 20-year climatology of nocturnal convection initiation over the central and southern Great Plains during the warm season. Mon. Wea. Rev., 145, 1615–1639, https://doi.org/10.1175/MWR-D-16-0340.1.
Rhines, P. B., 2002: Rossby waves. Encyclopedia of Atmospheric Sciences, Elsevier, 1923–1939.
Rossby, C.-G., 1939: Relation between variations in the intensity of the zonal circulation of the atmosphere and the displacements of the semi-permanent centers of action. J. Mar. Res., 2, 38–55.
Rossby, C.-G., 1940: Planetary flow patterns in the atmosphere. Quart. J. Roy. Meteor. Soc., 66, 68–87.
Rysman, J.-F., A. Lahellec, E. Vignon, C. Genthon, and S. Verrier, 2016: Characterization of atmospheric Ekman spirals at Dome C, Antarctica. Bound.-Layer Meteor., 160, 363–373, https://doi.org/10.1007/s10546-016-0144-y.
Saffin, L., J. Methven, J. Bland, B. Harvey, and C. Sanchez, 2021: Circulation conservation in the outflow of warm conveyor belts and consequences for Rossby wave evolution. Quart. J. Roy. Meteor. Soc., 147, 3587–3610, https://doi.org/10.1002/qj.4143.
Shaffer, G., 1972: A theory of time-dependent upwelling induced by a spatially- and temporally-varying wind with emphasis on the effects of a seabreeze–landbreeze cycle. Kiel. Meeresforsch., 28, 139–161.
Shapiro, A., E. Fedorovich, and S. Rahimi, 2016: A unified theory for the Great Plains nocturnal low-level jet. J. Atmos. Sci., 73, 3037–3057, https://doi.org/10.1175/JAS-D-15-0307.1.
Shapiro, A., E. Fedorovich, and J. G. Gebauer, 2018: Mesoscale ascent in nocturnal low-level jets. J. Atmos. Sci., 75, 1403–1427, https://doi.org/10.1175/JAS-D-17-0279.1.
Shapiro, A., J. G. Gebauer, and D. B. Parsons, 2022: Emergence of a nocturnal low-level jet from a broad baroclinic zone. J. Atmos. Sci., 79, 1363–1383, https://doi.org/10.1175/JAS-D-21-0187.1.
Sharan, M., and S. G. Gopalakrishnan, 1997: Comparative evaluation of eddy exchange coefficients for strong and weak wind stable boundary layer modelling. J. Appl. Meteor., 36, 545–559, https://doi.org/10.1175/1520-0450(1997)036<0545:CEOEEC>2.0.CO;2.
Shibuya, R., K. Sato, and M. Nakanishi, 2014: Diurnal wind cycles forcing inertial oscillations: A latitude-dependent resonance phenomenon. J. Atmos. Sci., 71, 767–781, https://doi.org/10.1175/JAS-D-13-0124.1.
Simpson, J. H., P. Hyder, T. P. Rippeth, and I. M. Lucas, 2002: Forced oscillations near the critical latitude for diurnal-inertial resonance. J. Phys. Oceanogr., 32, 177–187, https://doi.org/10.1175/1520-0485(2002)032<0177:FONTCL>2.0.CO;2.
Song, J., K. Liao, R. L. Coulter, and B. M. Lesht, 2005: Climatology of the low-level jet at the southern Great Plains atmospheric boundary layer experiments site. J. Appl. Meteor., 44, 1593–1606, https://doi.org/10.1175/JAM2294.1.
Stockwell, R. G., W. G. Large, and R. F. Milliff, 2004: Resonant inertial oscillations in moored buoy ocean surface winds. Tellus, 56A, 536–547, https://doi.org/10.3402/tellusa.v56i5.14478.
Stull, R. B., 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic, 666 pp.
Tan, Z.-M., and M. M. Farahani, 1998: An analytical study of the diurnal variations of wind in a semi-geostrophic Ekman boundary layer model. Bound.-Layer Meteor., 86, 313–332, https://doi.org/10.1023/A:1000694732459.
Tombrou, M., A. Dandou, C. Helmis, E. Akylas, G. Angelopoulos, H. Flocas, V. Assimakopoulos, and N. Soulakellis, 2007: Model evaluation of the atmospheric boundary layer and mixed-layer evolution. Bound.-Layer Meteor., 124, 61–79, https://doi.org/10.1007/s10546-006-9146-5.
Trier, S. B., and D. B. Parsons, 1993: Evolution of environmental conditions preceding the development of a nocturnal mesoscale convective complex. Mon. Wea. Rev., 121, 1078–1098, https://doi.org/10.1175/1520-0493(1993)121<1078:EOECPT>2.0.CO;2.
Trier, S. B., J. W. Wilson, D. A. Ahijevych, and R. A. Sobash, 2017: Mesoscale vertical motions near nocturnal convection initiation in PECAN. Mon. Wea. Rev., 145, 2919–2941, https://doi.org/10.1175/MWR-D-17-0005.1.
Tuttle, J. D., and C. A. Davis, 2006: Corridors of warm-season precipitation in the central United States. Mon. Wea. Rev., 134, 2297–2317, https://doi.org/10.1175/MWR3188.1.
Tyson, P. D., 1968: Velocity fluctuations in the mountain wind. J. Atmos. Sci., 25, 381–384, https://doi.org/10.1175/1520-0469(1968)025<0381:VFITMW>2.0.CO;2.
Vincze, M., N. Fenyvesi, M. Klein, J. Sommeria, S. Viboud, and Y. Ashkenazy, 2019: Evidence for wind-induced Ekman layer resonance based on rotating tank experiments. Europhys. Lett., 125, 44001, https://doi.org/10.1209/0295-5075/125/44001.
Wallace, J. M., 1975: Diurnal variations in precipitation and thunderstorm frequency over the conterminous United States. Mon. Wea. Rev., 103, 406–419, https://doi.org/10.1175/1520-0493(1975)103<0406:DVIPAT>2.0.CO;2.
Walters, C. K., and J. A. Winkler, 2001: Airflow configurations of warm season southerly low-level wind maxima in the Great Plains. Part I: Spatial and temporal characteristics and relationship to convection. Wea. Forecasting, 16, 513–530, https://doi.org/10.1175/1520-0434(2001)016<0513:ACOWSS>2.0.CO;2.
Walters, C. K., J. A. Winkler, R. P. Shadboldt, J. van Ravensway, and G. D. Bierly, 2008: A long-term climatology of southerly and northerly low-level jets for the central United States. Ann. Assoc. Amer. Geogr., 98, 521–552, https://doi.org/10.1080/00045600802046387.
Watanabe, M., and T. Hibiya, 2002: Global estimates of the wind-induced energy flux to inertial motions in the surface mixed layer. Geophys. Res. Lett., 29, 1239, https://doi.org/10.1029/2001GL014422.
Weckwerth, T. M., and D. B. Parsons, 2006: A review of convection initiation and motivation for IHOP_2002. Mon. Wea. Rev., 134, 5–22, https://doi.org/10.1175/MWR3067.1.
Weckwerth, T. M., J. Hanesiak, J. W. Wilson, S. B. Trier, S. K. Degelia, W. A. Gallus Jr., R. D. Roberts, and X. Wang, 2019: Nocturnal convection initiation during PECAN 2015. Bull. Amer. Meteor. Soc., 100, 2223–2239, https://doi.org/10.1175/BAMS-D-18-0299.1.
Weller, R. A., 1981: Observations of the velocity response to wind forcing in the upper ocean. J. Geophys. Res., 86, 1969–1977, https://doi.org/10.1029/JC086iC03p01969.
Weller, R. A., 1982: The relation of near-inertial motions observed in the mixed layer during the JASIN (1978) experiment to the local wind stress and to the quasi-geostrophic flow field. J. Phys. Oceanogr., 12, 1122–1136, https://doi.org/10.1175/1520-0485(1982)012<1122:TRONIM>2.0.CO;2.
Whiteman, C. D., X. Bian, and S. Zhong, 1997: Low-level jet climatology from enhanced rawinsonde observations at a site in the southern Great Plains. J. Appl. Meteor., 36, 1363–1376, https://doi.org/10.1175/1520-0450(1997)036<1363:LLJCFE>2.0.CO;2.
Whitt, D. B., and L. N. Thomas, 2015: Resonant generation and energetics of wind-forced near-inertial motions in a geostrophic flow. J. Phys. Oceanogr., 45, 181–208, https://doi.org/10.1175/JPO-D-14-0168.1.
Wilson, J. W., and R. D. Roberts, 2006: Summary of convective storm initiation and evolution during IHOP: Observational and modeling perspective. Mon. Wea. Rev., 134, 23–47, https://doi.org/10.1175/MWR3069.1.
Wirth, V., M. Riemer, E. K. M. Chang, and O. Martius, 2018: Rossby wave packets on the midlatitude waveguide—A review. Mon. Wea. Rev., 146, 1965–2001, https://doi.org/10.1175/MWR-D-16-0483.1.
Yamada, T., and G. Mellor, 1975: A simulation of the Wangara atmospheric boundary layer data. J. Atmos. Sci., 32, 2309–2329, https://doi.org/10.1175/1520-0469(1975)032<2309:ASOTWA>2.0.CO;2.
Zhang, X., S. F. DiMarco, D. C. Smith IV, M. K. Howard, A. E. Jochens, and R. D. Hetland, 2009: Near-resonant ocean response to sea breeze on a stratified continental shelf. J. Phys. Oceanogr., 39, 2137–2155, https://doi.org/10.1175/2009JPO4054.1.
Zhang, X., D. C. Smith IV, S. F. DiMarco, and R. D. Hetland, 2010: A numerical study of sea-breeze-driven ocean Poincare wave propagation and mixing near the critical latitude. J. Phys. Oceanogr., 40, 48–66, https://doi.org/10.1175/2009JPO4216.1.
Zhong, S., J. D. Fast, and X. Bian, 1996: A case study of the Great Plains low-level jet using wind profiler network data and a high-resolution numerical model. Mon. Wea. Rev., 124, 785–806, https://doi.org/10.1175/1520-0493(1996)124<0785:ACSOTG>2.0.CO;2.