Sensitivity of Idealized Moist Baroclinic Waves to Environmental Temperature and Moisture Content

D. J. Kirshbaum McGill University, Montreal, Quebec, Canada

Search for other papers by D. J. Kirshbaum in
Current site
Google Scholar
PubMed
Close
,
T. M. Merlis McGill University, Montreal, Quebec, Canada

Search for other papers by T. M. Merlis in
Current site
Google Scholar
PubMed
Close
,
J. R. Gyakum McGill University, Montreal, Quebec, Canada

Search for other papers by J. R. Gyakum in
Current site
Google Scholar
PubMed
Close
, and
R. McTaggart-Cowan Environment and Climate Change Canada, Dorval, Quebec, Canada

Search for other papers by R. McTaggart-Cowan in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Idealized simulations are used to examine the sensitivity of moist baroclinic wave growth to environmental temperature and moisture content. With relative humidity held fixed, the surface temperature at 45°N, denoted T0, is varied from 275 to 290 K. As T0 increases, the atmospheric moisture content, moist instability, and moist available potential energy also increase. For the chosen initial configuration, moist waves develop larger eddy kinetic energy Ke than corresponding dry waves, but enhanced diabatic heating at larger T0 does not further increase Ke. This finding is linked to a warm-frontal cyclonic potential vorticity (PV) anomaly that strengthens and shifts downstream at larger T0 owing to increased diabatic heating along the frontal cloud band. This eastward shift feeds back negatively on the parent cyclone by increasing the downstream export of mechanical energy aloft and degrading the phasing between dry baroclinic vertical motion and buoyancy within the warm sector. The latter suppresses the conversion from eddy potential energy to Ke [C(Pe, Ke)], offsetting a direct enhancement of C(Pe, Ke) by diabatic heating. Compared to their dry counterparts, isolated moist waves (initiated by a single finite-amplitude PV anomaly) display a similar sensitivity to T0, while periodic wave trains (initiated by multiple such anomalies) exhibit a stronger negative relationship. The latter stems from anticyclonic diabatic PV anomalies aloft that originate along the warm front and recirculate through the system to interact with the upper-level trough. This interaction leads to a horizontal forward wave tilt that enhances the conversion of wave Ke into zonal-mean kinetic energy.

© 2018 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: D. J. Kirshbaum, daniel.kirshbaum@mcgill.ca

Abstract

Idealized simulations are used to examine the sensitivity of moist baroclinic wave growth to environmental temperature and moisture content. With relative humidity held fixed, the surface temperature at 45°N, denoted T0, is varied from 275 to 290 K. As T0 increases, the atmospheric moisture content, moist instability, and moist available potential energy also increase. For the chosen initial configuration, moist waves develop larger eddy kinetic energy Ke than corresponding dry waves, but enhanced diabatic heating at larger T0 does not further increase Ke. This finding is linked to a warm-frontal cyclonic potential vorticity (PV) anomaly that strengthens and shifts downstream at larger T0 owing to increased diabatic heating along the frontal cloud band. This eastward shift feeds back negatively on the parent cyclone by increasing the downstream export of mechanical energy aloft and degrading the phasing between dry baroclinic vertical motion and buoyancy within the warm sector. The latter suppresses the conversion from eddy potential energy to Ke [C(Pe, Ke)], offsetting a direct enhancement of C(Pe, Ke) by diabatic heating. Compared to their dry counterparts, isolated moist waves (initiated by a single finite-amplitude PV anomaly) display a similar sensitivity to T0, while periodic wave trains (initiated by multiple such anomalies) exhibit a stronger negative relationship. The latter stems from anticyclonic diabatic PV anomalies aloft that originate along the warm front and recirculate through the system to interact with the upper-level trough. This interaction leads to a horizontal forward wave tilt that enhances the conversion of wave Ke into zonal-mean kinetic energy.

© 2018 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: D. J. Kirshbaum, daniel.kirshbaum@mcgill.ca

1. Introduction

Midlatitude cyclones form during all four seasons, and over a wide range of latitudes, in both maritime and continental regions. Thus, the mean environmental temperatures in which they develop (here denoted T0) can vary widely. Moreover, owing to anthropogenic emissions of greenhouse gases, the distributions of T0 experienced by these cyclones may gradually change in coming decades (IPCC 2013), although likely to a smaller degree than seasonal and latitudinal variations in the current climate. To interpret the behavior of such cyclones over different seasons, latitudes, and climates, it is useful to conceptually understand their sensitivities to T0. This sensitivity has not received extensive attention from a dynamical perspective, in part because it is secondary in classic quasigeostrophic theories of dry baroclinic instability (Eady 1949; Charney 1947). In these theories, the linear growth rate of baroclinic instabilities depends on the meridional baroclinicity and static stability, neither of which depends strongly on T0.

Midlatitude cyclones may still be regulated by T0 through its indirect control over diabatic processes like latent heat release. At a fixed relative humidity, the atmospheric water vapor content increases by around 7% K−1 and, hence, nearly doubles with every 10 K of warming. This increased moisture content supports enhanced latent heat release, which influences the parent baroclinic wave by destabilizing ascending motions and generating diabatic potential vorticity (PV) anomalies that interact with dry baroclinic PV anomalies (e.g., Hoskins et al. 1985; Davis and Emanuel 1991; De Vries et al. 2010). In the vertical, a diabatically generated PV tendency dipole straddles the active cloud layer, with cyclonic tendencies below and anticyclonic tendencies above.

Using the semigeostrophic equations with parameterized condensational heating, Emanuel et al. (1987) found that such heating led to faster wave growth, at a reduced horizontal scale. However, in subsequent studies that modified the parameterization of cloud latent heating for added realism, the positive impacts on cyclone growth weakened or even reversed (Whitaker and Davis 1994; Mak 1994). Lapeyre and Held (2004) evaluated the dynamics of moist baroclinic eddies using a two-layer quasigeostrophic model, also with parameterized cloud diabatic heating. As the heating rate was increased, the eddies transitioned from baroclinic wave structures analogous to dry cyclones to intense, vortex-dominated structures with no dry analog.

Case studies mostly support Emanuel et al. (1987)’s finding that latent heat release contributes to parent-cyclone intensification. For example, using real-case numerical simulations and piecewise PV inversion, Stoelinga (1996) found that diabatically generated PV anomalies within an intense Atlantic cyclone greatly amplified the low-level cyclonic circulation and the phase locking between lower- and upper-level PV anomalies. Similarly, Ahmadi-Givi et al. (2004) found that a diabatically generated low-level cyclonic PV anomaly dominated the spinup of an Atlantic cyclone at low levels. The cyclone growth was aided by the corresponding anticyclonic PV anomaly aloft, which eroded the downstream flank of the upper trough to reinforce an upstream phase tilt. While other case studies have reached similar conclusions (e.g., Reed et al. 1988; Willison et al. 2013), some have also found the opposite. Notably, in a case study of a winter cyclone over the central United States, Davis (1992) found that differential eastward propagation of the upper-level dry baroclinic PV anomaly and a lower-level diabatic PV anomaly led to cyclolysis.

Further insights into the dynamics and sensitivities of moist baroclinic waves have been gained from idealized initial-value experiments in periodic limited-area domains. Such experiments should be interpreted with caution, because their prescribed domain size constrains the wave dynamics and their periodicity promotes unphysical self-interactions, as disturbances shed downwind may recirculate to the upstream side. Nonetheless, most such simulations suggest an amplification of the waves due to moist processes (e.g., Whitaker and Davis 1994; Balasubramanian and Yau 1994; Fantini 2004; Boutle et al. 2011; Booth et al. 2013). Only a few such studies have explored the sensitivities of the waves to T0, including Whitaker and Davis (1994) and Boutle et al. (2011, hereafter BBP11), who both found a modest but systematic wave intensification in warmer atmospheres. Because these findings reinforced the expectation that stronger diabatic heating intensifies baroclinic waves, they were not analyzed in detail. Importantly, given the small sampling of initial jets, initial wave perturbations, and other factors in these studies, their findings may not generalize to all baroclinic waves.

By contrast, climate studies have shown mixed sensitivities of cyclone intensity to T0 (e.g., O’Gorman 2011; Pfahl et al. 2015; Büehler and Pfahl 2017). Using an idealized global climate model (GCM), Pfahl et al. (2015) found that, on average, these cyclones tend to amplify with increasing T0 in cooler atmospheres but weaken with increasing T0 in warmer atmospheres. This nonmonotonic sensitivity tracked that of the zonal-mean available potential energy Pm, implying that Pm serves to broadly regulate it. However, the strongest 10% of cyclones defied this trend by exhibiting a nearly monotonic intensification as T0 was increased. In GCM climate projections, no consensus has been reached on whether midlatitude cyclones will intensify in warmer future climates (e.g., Bengtsson et al. 2009; Catto et al. 2011; Zappa et al. 2013; Colle et al. 2013). Also, using the “pseudo global warming” approach of Schär et al. (1996) over the East Coast of the United States, Marciano et al. (2015) found that most, but not all, simulated cyclones strengthened when the environment was warmed in a fashion consistent with GCM projections.

Given the contrasting sensitivities of midlatitude cyclones to T0 found previously, and the possibility that the nature of this sensitivity may depend on numerous factors, this problem remains unresolved and merits further investigation. Of particular interest herein are the dynamical feedbacks of enhanced diabatic PV anomalies in warmer atmospheres on their parent waves. To address this question, we conduct idealized initial-value experiments of moist baroclinic waves, including both isolated waves in an otherwise undisturbed flow and periodic wave trains, where T0 is systematically varied. Section 2 describes the numerical configuration, with the isolated- and periodic-wave simulations presented in sections 3 and 4, respectively. Further discussion is provided in section 5, followed by the conclusions in section 6.

2. Model setup

a. Model configuration

The Advanced Research version of the Weather Research and Forecasting (WRF) Model, version 3.7 (Skamarock et al. 2008), an Eulerian finite-difference model that integrates the nonhydrostatic and nonlinear atmospheric equations using third-order Runge–Kutta time differencing and a split time step for stability of sound waves, is used for the numerical experiments. All aspects of the simulation design are geared toward isolating the problem of interest in the simplest possible framework. Thus, some physical processes that may impact baroclinic waves, but do not bear directly on our main scientific question, are deliberately omitted and deferred to future work.

A single domain is used with dimensions of Lx = 16 000 km (zonal) by Ly = 12 000 km (meridional) by Lz = 20 km (vertical). The nominal horizontal grid spacing is Δh = 100 km, which is relatively coarse by modern numerical weather prediction (NWP) standards but still sufficient to capture the key sensitivities of the waves to T0 (as found by selected experiments reducing Δh down to 25 km; not shown). A stretched hydrostatic-mass vertical coordinate with 101 unstaggered levels is used, with a grid spacing ranging from ~90 m near the surface to ~400 m at the model top. The uppermost 4 km uses a Rayleigh damping layer to minimize wave reflections off the rigid model top. The lateral boundaries are periodic in x and rigid in y. In keeping with previous initial-value experiments in limited-area domains, an f-plane approximation is used with f = f0 = 2Ωsinϕ0, where Ω is Earth’s angular velocity and ϕ0 = 45°N. Because WRF uses a Cartesian horizontal grid, its neglect of spherical metric terms—which promotes cyclonic over anticyclonic Rossby wave breaking (e.g., Govindasamy and Garner 1997)—may limit the real-world applicability of the experimental findings.

Default physical parameterizations include the WRF single-moment 6-class cloud microphysics scheme (Hong and Lim 2006), the Kain–Fritsch cumulus parameterization scheme (Kain 2004), the Yonsei University planetary boundary layer scheme with Monin–Obukhov similarity theory in the surface layer (Hong et al. 2006), and horizontal mixing along coordinate surfaces using a Smagorinsky–Lilly scheme. The lower boundary is a no-slip land surface with zero sensible and latent heat fluxes. Thus, flows that are initially dry stay dry over time. This simplification, as well as the neglect of both short- and longwave radiation, is enforced to ensure that the vast majority of diabatic heating is owing to phase changes of water substance. While these simplifications limit the realism of the simulations, they also facilitate physical understanding.

b. Initialization

The model is initialized with a baroclinic jet in thermal wind balance, following the three-step procedure proposed in Olson and Colle (2007) and adopted by Schemm et al. (2013) with some minor modifications. First, the pressure field at a height of zref = 4 km is specified as
e1
e2
where yc = Ly/2, P0 = 607 hPa, P1 = 1200 km, and P2 = 1800 km are small- and large-scale pressure “packings,” which control pressure gradients near the domain center, and Ay1 = 12 and Ay1 = 17 hPa are the corresponding small- and large-scale pressure-gradient amplitudes.
Second, the yz virtual temperature distribution is specified as follows:
e3
where Z = Z(y) is the tropopause height, By1 = 10 K, By2 = 12 K, and Tυ0 = T0(1 + 0.61qυ0) is reference surface virtual temperature at yc (corresponding to a latitude of 45°N), with T0 the surface temperature and qυ0 the surface water vapor mixing ratio. The virtual temperature lapse rates for the troposphere Γt and stratosphere Γs are
e4
e5
where γt = 6.0 K km−1 and γs = 0.3 K km−1. For simplicity, the initial relative humidity profile RH(z) is horizontally homogeneous and specified by a piecewise-linear function
e6
where γRH = 0.05 km−1, RH1 is the surface value, and RH2 is the minimum value aloft. Note that, unlike previous studies using this initialization technique, we specify Tυ rather than T because it ties more directly to air density ρ, geopotential height, and the thermal wind.
Finally, the full pressure distribution is obtained through hydrostatic vertical integration upward and downward from zref to the domain top and bottom, respectively. The tropopause height is
e7
where
e8
Z0 = 8.5 km, and Zdiff ≈ 3.5 km is half of the total increase in Z from the pole to the tropics. An example baroclinic jet for T0 = 290 K, RH1 = 0.8, and RH2 = 0.1 is shown in Fig. 1a, with large differences in surface virtual potential temperature θυ (~40 K) and qυ (~24 g kg−1) between the poles and tropics, and a ~44 m s−1 jet maximum along the midlatitude tropopause.
Fig. 1.
Fig. 1.

Details of the model initialization for the T290-moist ISO case: (a) water vapor mixing ratio (filled contours), virtual potential temperature (thin lines, interval: 8 K), zonal winds (thick lines, interval: 6 m s−1), tropopause height (thick dashed line) of initial jet profile; (b) PV′ (filled) and meridional winds [m s−1; positive values (solid lines) indicate flow into the page and negative values (dashed lines) indicate flow out of the page].

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The initial jet is defined using the full moist equations and is thus perfectly balanced. However, because the surface winds are nonzero and the model lower boundary is no slip, a zonally symmetric Ekman layer develops over the first 24 h of model integration. Thereafter, the jet remains in a steady state if left unperturbed (not shown).

To create growing wave disturbances, an initial perturbation is added to the jet. While normal-mode perturbations can be defined for any baroclinically unstable environment (e.g., Simmons and Hoskins 1976; Merlis and Schneider 2009), such perturbations do not generally explain transient cyclogenesis (Farrell 1985). Modal calculations with a saturated atmosphere can capture some of the destabilizing effects of latent heat release (e.g., Emanuel et al. 1987) but have collapsing updraft area for warm basic states that are not reproduced in finite-amplitude baroclinic life cycle simulations (O’Gorman et al. 2017). Last, optimal transient perturbations can also be obtained (Farrell 1989), but their derivation assumes zonal periodicity and does not naturally account for nonlinear diabatic heating.

We instead prescribe an isolated, finite-amplitude upper-level PV anomaly centered above the low-level baroclinic zone to represent Petterssen’s “type B” cyclogenesis configuration (Petterssen and Smebye 1971). Pioneered by Takayabu (1991) and Schär and Wernli (1993), the use of such isolated, finite-amplitude perturbations avoids some of the complexities associated with periodic moist waves (as will be seen section 4). As in Schemm et al. (2013), we add an Ertel PV perturbation of the form
e9
to the initial state, where xp, yp = yc, and zp = 8 km are the central coordinates, Dx = Dy = 500 km, Dz = 4 km, and PV0 = 2 PVU (where 1 PVU = 10−6 m2 s−1 K kg−1). This perturbation is converted to quasigeostrophic (QG) potential vorticity q′ through the relation and then inverted into geopotential perturbation ϕ′ using a simple QG inversion technique. The resulting geostrophic wind perturbation for the case of T0 = 290 K, RH1 = 0.8, and RH2 = 0.1 is shown in Fig. 1b, depicting a balanced cyclonic vortex surrounding the core of the PV anomaly.

Eight initial baroclinic jets are defined using different combinations of T0, RH1, and RH2, the details of which are summarized in Table 1. These cases span a range of T0 (275–290 K, in 5-K increments) that roughly represents the variation in T0 between the midlatitude winter and spring/autumn seasons. They also incorporate two RH profiles, one moist with RH1 = 0.8 and RH2 = 0.1 and the other dry with RH1 = RH2 = 0. By using fixed meridional and vertical Tυ gradients and tropopause heights, the experiments neglect changes in midlatitude storm tracks that might accompany global climate change (e.g., Schneider et al. 2010). Thus, some factors that may affect baroclinic waves in future climates are omitted.

Table 1.

Details of the eight simulations with varying T0 and RH, including prescriptions of varying parameters (all of which are defined in the text).

Table 1.

The simulations are named according to their T0 and RH; for example, the moist case with T0 = 290 K in Fig. 1a is named T290-moist (Table 1). Two sets of eight simulations are considered, each set with different initial perturbations to initiate baroclinic waves. The first (“ISO”) uses an isolated PV perturbation centered at xp = xc, where xc = Lx/2. To help bridge the gap between the ISO simulations and the periodic waves simulated in related studies (Whitaker and Davis 1994; BBP11), the second (“PER”) superposes three of the same PV perturbations, equally spaced over the zonal domain, to initiate a periodic wave train. The spacing between these perturbations (5333 km) is set to roughly match the natural zonal wavelengths that develop in the ISO simulations (~5400 km, as will be seen).

The tropospheric-mean Brunt–Väisälä frequency of Ntrop ≈ 0.012 s−1 varies little between the simulations, owing to their fixed Tυ lapse rates (Table 1). Similarly, the maximum jet speed (Umax ≈ 44 m s−1) is invariant because the meridional Tυ gradients are fixed. However, the zonal-mean available potential energy Pm [calculated for both dry and moist flows using the Munkres algorithm, as detailed in Stansifer et al. (2017)] decreases from 3.9 to 3.5 MJ m−2 in the dry cases over the range of T0. This decrease can be explained by the approximate Pm equation of Lorenz (1955):
e10
where Γ = −δzT, Γd = g/cp, ps, and pt are the surface and model-top pressures, angle brackets denote domainwide horizontal averages on pressure surfaces, and asterisks denote perturbations from those averages. As T0 is increased from 275 to 290 K, 〈T−1 decreases by 10%, as does Pm.

The value of Pm of the moist simulations exceeds that in the dry simulations and also exhibits the opposite sensitivity to T0: rather than decreasing over the range of T0, it increases from 4.3 to 5.3 MJ m−2. This trend is owing to increased latent heat release, and consequent static destabilization, in the process of rearranging air parcels to obtain the reference state for the Pm calculation. Also, the maximum initial convective available potential energy at the southern end of the domain (CAPEmax) increases dramatically with T0, from 112 J kg−1 (T275-moist) to 3180 J kg−1 (T290-moist), with nonzero values shifting northward from around 2000 to 500 km equatorward of yc (not shown). Except for a similar northward shift in nonzero values at larger T0, the convective inhibition (CIN) exhibits minimal sensitivity to T0, with mean values of around 50 J kg−1 in areas of nonzero CAPE (not shown).

3. Isolated waves

a. Overview

Snapshots of surface equivalent potential temperature θe, cloud hydrometeor path, surface pressure, and surface winds for four ISO simulations (T275-dry, T290-dry, T275-moist, and T290-moist) are shown in Fig. 2. These are taken at 0 h on day 6, within 24 h of the cyclones reaching their minimum surface pressure. Over this range of parameter space, the simulated waves exhibit only subtle qualitative differences—notably that the magnitude of the surface low increases slightly in the moist cyclones and decreases slightly as T0 is increased (Figs. 2b,d). The most noticeable structural difference between the two moist simulations is an eastward extension of the warm-frontal cloud band in T290-moist. The colder simulations also exhibit larger ps, which follows directly from (1) and (2): with a fixed pressure distribution at zref = 4 km across all simulations, the mean ps increases with the density of the underlying layer, which is larger in the colder flows.

Fig. 2.
Fig. 2.

Snapshots of surface (filled color contours), pressure (black lines, interval: 10 hPa), and wind vectors (reference vector in top-right corner of each panel), along with cloud hydrometeor path (filled gray contours), at 0 h on day 6 for four ISO simulations. Only an 8000-km-wide region centered on the near-surface pressure minimum (at x = xc) is shown. The abscissa represents distance from the surface cyclone center.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Quantitative differences in simulated wave growth are evaluated within a cyclone-centered reference frame, which occupies the full domain in y and z but is limited to a zonal length of lx = 5400 km to roughly encompass one zonal wavelength (e.g., Fig. 2). The cyclone position is tracked using an algorithm that steps chronologically through each model output time, calculating (ΔSLP)min, the minimum surface pressure perturbation relative to the zonally averaged initial value, and the maximum horizontal Laplacian of surface pressure, over the entire domain. Once these two extrema are zonally separated by less than lx/2 and (ΔSLP)min ≤ −5 hPa, the cyclone center (xc, yc) is assigned to the location of (ΔSLP)min. The algorithm then tracks this feature at subsequent times by restricting the search area to x ± lx/2 of the previous cyclone position, thus preventing it from jumping to developing cyclones elsewhere in the domain.

As the ISO cyclones mature, they undergo a cyclonic wrapping process that ultimately results in the base of the upper-level trough impinging upon the downstream boundary of the control volume by around day 8. Nonetheless, analyses with the reference frame shifted eastward by up to 1000 km yielded similar results to those for a cyclone-centered frame. Hence, the analyses below (which focus on the cyclone growth over days 1–7) are not strongly sensitive to the precise reference-frame position.

Because of the increased atmospheric moisture content in warmer environments, the cumulative precipitation within the analysis region 〈Rtot〉 increases monotonically with T0 in the moist ISO simulations and more than doubles over the range of T0 (Table 2). This trend stems primarily from increased parameterized convective precipitation 〈Rcu〉 driven by the rapid increase in moist instability at larger T0 (Table 1). By contrast, the explicit precipitation from the cloud-microphysics scheme 〈Rmic〉 increases by only ~50% over the range of T0.

Table 2.

Horizontally averaged cumulative precipitation (mm) at 0 h on day 10 from the cloud microphysics 〈Rmic〉 and cumulus 〈Rcu〉 schemes, along with the sum of the two 〈Rtot〉, for all moist simulations. All averages are taken over the limited-area cyclone-tracking analysis domain.

Table 2.

To evaluate the sensitivity of the wave dynamics to T0, Fig. 3 compares time series of three diagnostics, the first of which is the eddy kinetic energy,
e11
where u′ and υ′ are the perturbation zonal and meridional winds. An overbar [e.g., ] denotes a zonal average over the cyclone-centered analysis area and the prime [e.g., u′ = u′(x, y, z, t)] denotes a deviation from that average. To evaluate (11), the model data are interpolated from their native coordinates to p coordinates with bounds of 1050 and 25 hPa. To avoid errors incurred by vertical extrapolation, data are masked where these bounds fall outside the range of simulated pressures in each column. The second quantity, the “conversion efficiency,” normalizes Ke by initial Pm to indicate the fraction of available potential energy that is converted into wave motions. The third quantity, (ΔSLP)min, provides a complementary indicator of cyclone intensity.
Fig. 3.
Fig. 3.

Time series of (a),(b) Ke, (c),(d) conversion efficiency (the ratio of Ke to initial Pm), and (e),(f) for the (left) dry and (right) moist ISO simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

In the dry ISO simulations, Ke of the tracked wave increases until about day 7 and then decays (Fig. 3a). Averaged over the growth phase, Ke and |(ΔSLP)min| decrease by ~15% as T0 is increased from 275 to 290 K. This negative sensitivity stems, at least in part, from the negative sensitivity of Pm to T0 (Table 1). Thus, when Ke is normalized by Pm to give the conversion efficiency, the gap between the different simulations is narrowed (Fig. 3c). An additional contribution to this negative sensitivity is the decreased air density in the warmer cases, which causes the tropospheric mass, and hence the wave Ke, to decrease by ~5% over the range of T0.

Compared to their dry counterparts, the moist ISO waves intensify faster and exhibit less sensitivity to T0, during their early growth phase (days 2–3.5; Figs. 3b,f). This amplification stems from cloud diabatic heating, which commences early on day 2 (as will be shown). During this period, the reduced dry growth of the warmer waves (as evident from Figs. 3a and 3e) is offset by increased diabatic heating to leave minimal residual sensitivity to T0.

The moist waves reach maturity over days 3.5–7, attaining a ~30% larger peak Ke and a ~10-hPa smaller (ΔSLP)min than the corresponding dry ISO waves. Thus, cloud diabatic heating clearly intensifies the parent cyclone. However, consistent with Schemm et al. (2013), the degree of diabatic enhancement for this configuration is small compared to that in some other studies (e.g., BBP11; Stoelinga 1996). During this phase of development, the warmer waves intensify less rapidly than the colder waves, leading to a monotonic decrease in maximum Ke and |(ΔSLP)min| with increasing T0 (Figs. 3b,f). Although the percentage decrease in Ke over the simulated range of T0 is similar to that in the corresponding dry cases, the conversion efficiency declines much faster with increasing T0 (Figs. 3c,d). Thus, as the environment warms, the moist cyclones become increasingly inefficient at converting environmental Pm into wave Ke.

b. Energetics

To interpret the sensitivities of the simulated wave growth to T0, we analyze the Ke budget. Following a similar derivation as Orlanski and Katzfey (1991), the time tendency of ke (Ke per unit mass) may be written in pressure coordinates as
e12
where is the full 3D wind vector minus the zonal speed of the reference frame , v is the 2D horizontal wind vector, is the pressure velocity, and F is the frictional acceleration. Because ω is not a state variable in WRF, it is estimated using the hydrostatic approximation , where ρ is density and w is vertical velocity. The terms on the right-hand side (rhs) of (12), which represent energy transfers per unit mass, include (i) advection of ke (“adv”), (ii) ageostrophic geopotential fluxes [“ageo,” after Orlanski and Katzfey (1991)], (iii) baroclinic conversion from eddy potential energy pe to ke , (iv) conversion from ke to zonal-mean kinetic energy , and frictional dissipation . We evaluate the specific-energy terms in (12) (denoted by lowercase letters) at each grid point, then integrate over mass using the operator in (11) to obtain the full terms (denoted by uppercase letters).

Consistent with the Ke time series in Fig. 3a, the Ke tendency of the dry ISO simulations is positive until ~day 7 and weakens slightly at larger T0 (Fig. 4a). The latter effect is driven by reduced C(Pe, Ke) and yields a consequent weakening of and (Figs. 4b–d). This explanation is supported by time series of the differences in C(Pe, Ke), , and between these two simulations in Fig. 5. While differences in C(Pe, Ke) appear immediately, differences in and do not emerge until days 2–3, once clear contrasts in Ke are already established (Fig. 3a). Thus, the sensitivity of C(Pe, Ke) to T0 appears to drive differences in system evolution that ultimately affect the other budget terms. Finally, the two boundary flux terms (ADV and AGEO) are negative but similar in both cases (Figs. 4e,f), indicating a net export of mechanical energy out of the analysis volume.

Fig. 4.
Fig. 4.

Time series of Ke budget terms from (12), for the dry ISO simulations. Both the rate of change of the model-diagnosed Ke and the sum of the Ke-budget source terms [the lhs and rhs of (12), respectively] are respectively shown by solid and dotted lines in (a).

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Fig. 5.
Fig. 5.

Time series of the differences of three terms of the Ke budget between the T275-dry and T290-dry ISO simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The moist cases develop larger C(Pe, Ke) and Ke tendencies than the corresponding dry cases, and these terms do not noticeably depend on T0 until the onset of the maturation phase (day 3.5; Figs. 6a,b). Thereafter, the Ke tendency weakens with increasing T0. Unlike in the dry cases, this decrease is not driven by C(Pe, Ke); it instead stems from a more negative AGEO term (Fig. 6e), related to the aforementioned structural differences in the warm-frontal cloud band (Figs. 2c,d). As this cloud band shifts eastward at larger T0, so does the low-level warm advection, which approaches the downwind boundary of the analysis volume (Fig. 7). Because both the diabatic heating and the warm advection contribute to upper-level ridge building and divergence (e.g., Holton 1992), the ageostrophic geopotential fluxes across the downwind boundary increase.

Fig. 6.
Fig. 6.

As in Fig. 4, but for the moist ISO simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Fig. 7.
Fig. 7.

Mean 200–400-hPa geopotential height perturbation (filled) and ageostrophic wind vectors, along with mean 100–1000-hPa diabatic heating rate (single white contour at 10−5 K s−1) and 700–900-hPa geostrophic temperature advection (black contours, 10−4 K s−1), at 0 h on day 6 of the ISO experiments. Dashed lines denote lateral boundaries of cyclone-following control volume, and an asterisk denotes the cyclone-center position.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The increased AGEO magnitude at larger T0 is partially offset by a corresponding reduction in the magnitudes of and (Figs. 6c,d), again owing to the weakening of the cyclone. After day 7, AGEO switches sign (Fig. 6e) as cyclonic wrapping causes the base of the trough–ridge couplet to rotate eastward. The easterly ageostrophic flow between the trough and ridge (visible to the south of the cyclone center on day 6 in Fig. 7) reaches the eastern boundary of the analysis volume to transport relatively high geopotential air into the analysis domain (not shown).

The fact that Ke does not increase with T0 in the maturing moist cyclones suggests that increased latent heat release, which is supported in warmer environments with larger moisture content, does not necessarily energize the baroclinic waves. If, for all else equal, increased latent heat release occurred in warm and ascending portions of the cyclone, it would directly enhance and indirectly enhance by destabilizing ascending motions. This combination would increase in (12) [and hence C(Pe, Ke)], which would tend to enhance Ke. While such an enhancement is indeed found between the dry and moist simulations at a given T0, it is absent in the moist simulations as T0 is increased (after day 3.5), despite a rapid increase in Pm. Thus, in these experiments, the dynamical feedbacks of enhanced diabatic heating at larger T0 act to inhibit, rather than enhance, the growth of the maturing cyclone.

c. QG ω analysis

To understand why the baroclinic conversion, the main process by which Ke is generated, does not increase with T0 in the moist simulations, we decompose it using the QG ω equation (e.g., Lackmann 2011),
e13
where is the geostrophic wind, , is the horizontally averaged θ profile over the full domain at time t, and J is the diabatic heating rate (including contributions from parameterized cloud microphysics, cumulus convection, and PBL mixing). The rhs contains the forcings for vertical motion, including, from left to right, (i) differential vorticity advection, (ii) the Laplacian of the thermal advection, and (iii) minus the Laplacian of the diabatic heating. Because the first two terms are not independent and exhibit a large degree of cancellation (e.g., Hoskins et al. 1978), we combine them into a single dry “baroclinic” term (term B). The third term is interpreted as the “diabatic” term (term D).
Domainwide inversion of the elliptical operator on the left-hand side (lhs) of (13) is performed by taking a Fourier transform of (13) in x and discretizing its derivatives in y and z using second-order finite differencing. For each Fourier mode k, a matrix is constructed relating , the Fourier transform of , to the rhs forcing terms. This matrix is inverted separately for each rhs forcing term i to obtain the corresponding component of , and an inverse Fourier transform is taken to retrieve . The contributions of each component of QG vertical motion to , the QG version of , are then obtained:
e14
e15
where and are the perturbation pressure velocities attributed to terms B and D in (13). In the following, we integrate the two rhs terms over mass using the operator in (11) and denote the resulting quantities terms B and D. Because (14) only decomposes the vertical motion and not , it only exposes the impacts of different sources of ω on the full field. A full decomposition, including a breakdown of into its different contributions, is omitted herein but represents a potentially interesting topic for future work.

Figures 8a and 8b compare C(Pe, Ke) diagnosed directly from the simulations against calculated from (14), for the dry and moist T275 and T290 simulations, respectively. Despite its simplifications, the QG analysis reasonably represents the magnitude, time evolution, and T0 sensitivity of C(Pe, Ke). While it fails to capture the very slight decrease in C(Pe, Ke) between the T275-moist and T290-moist simulations (Fig. 8b), it correctly indicates that this sensitivity is much weaker than that in the corresponding dry simulations.

Fig. 8.
Fig. 8.

QG analysis of C(Pe, Ke) in the ISO simulations. See section 3c for definitions of terms.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

In the dry cases, the diabatic component (term D) is negligible, so the dry baroclinic component (term B) fully accounts for the time evolution of (Fig. 8c). This term strengthens during the cyclone growth phase, then weakens during the cyclone dissipation phase, both at a generally weaker magnitude at larger T0. In the moist cases, terms B and D both increase modestly (Fig. 8d), causing a 20%–30% net increase in over that in the corresponding dry cases. During the early growth phase, an increase in term D gives rise to a slight increase in at larger T0. However, during the maturation phase, this increase is offset by a decrease in term B, leading to minimal net sensitivity of to T0 (Fig. 8d). These competing sensitivities largely explain the failure of the maturing moist waves to intensify at larger T0 despite their increased latent heat release.

To aid the physical interpretation of this finding, we use an approximate adiabatic form of (13) derived by Trenberth (1978):
e16
which relates QG vertical motion to geostrophic vorticity advection by the thermal wind. Figure 9 presents several quantities relevant to (16) at 0 h on day 5, from three simulations (T275-dry, T275-moist, and T290-moist) that span the range of simulated cloud diabatic heating rates.
Fig. 9.
Fig. 9.

Analysis of PV and baroclinic vertical motion in the (left) T275-dry, (center) T275-moist, and (right) T290-moist ISO simulations at 0 h on day 5. (a)–(c) Mean 100–1000-hPa diabatic heating rate (filled) along with mean 700–900-hPa Ertel PV (contours, PVU) and wind vectors. (d)–(f) Mean 700–900-hPa dry-baroclinic rhs forcing term of (16) [filled, with warm (cold) colors indicating forcing for ascent (descent)], geostrophic absolute vorticity (contours, 10−5 s−1), and thermal-wind vectors. (g)–(i) Mean 700–900-hPa (filled) and retrieved (contours, Pa s−1). For reference, (a)–(f) include a single contour of (red thick line) to outline the warm sector. The asterisk denotes the center of the surface cyclone.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

In the T275-dry case, a low-level mesoscale cyclonic PV anomaly, associated with frictional wind turning in baroclinic zones (e.g., Stoelinga 1996; Adamson et al. 2006), wraps cyclonically around the surface low (Fig. 9a). In the moist cases, this PV anomaly is enhanced by diabatic PV generation along the frontal cloud band and shifts progressively eastward as T0 is increased (Figs. 9b,c). This eastward shift becomes prominent during the transition from the early growth phase to the maturation phase (Fig. 10). On day 3, the frontal cloud bands in the T275-moist and T290-moist cases are similarly located, but increased diabatic PV generation in the latter extends the PV anomaly slightly eastward (Figs. 10a,b). As a result, the southerly branch of the induced cyclonic circulation, carrying high- air from the tropics, also shifts eastward. Owing to intense latent heat release and resulting diabatic PV generation along the nose of this ascending, conditionally unstable flow, the frontal PV maximum becomes phase locked to it, and both propagate eastward, in the T290-moist case (Figs. 10d and 10f). A similar process occurs in T275-moist (Figs. 10d and 10f), but reduced diabatic PV generation limits the eastward propagation.

Fig. 10.
Fig. 10.

Evolution over days 3–4 of mean 700–900-hPa PV (solid black lines, contours of 1, 1.5, and 2 PVU) and wind vectors, along with CAPE (based on lifting the parcel in each column with the maximum below 500 hPa; red fill) and mean 100–1000-hPa diabatic heating rate (gray fill) for the (a),(c),(e) T275-moist and (b),(d),(f) T290-moist simulations. A single contour of (red thick line) is also shown to outline the warm sector.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Similar to the PV, the low-level geostrophic absolute vorticity in the T275-dry simulation is maximized in a band wrapping around the low center (Fig. 9d). Advection of by the thermal wind generates a dipole in the rhs forcing term in (16), with forcing for descent along the southern flank of the maximum (just east of the surface low) and forcing for ascent along the northeastern flank. With the forcing for ascent lying entirely within the northern warm sector, and the forcing for descent lying mostly outside the warm sector, the phasing between the baroclinic forcing and is favorable for maximizing term B.

In the T275-moist simulation, diabatic PV generation in the warm-frontal cloud band increases and, hence, the local forcing for ascent just downwind (Fig. 9e). The eastward shift of the maximum along the warm front causes the rhs forcing for ascent to also migrate eastward toward the edge of the warm sector and the forcing for descent to migrate partially into the warm sector. This diabatically induced phase difference between rhs forcing and increases further in the T290-moist simulation, where the forcing for ascent crosses the northeastern edge of the warm sector as the forcing for descent becomes nearly fully immersed within the warm sector (Fig. 9f). Thus, while latent heat release increases the forcing for vertical motion, it also increases the phase difference between this forcing and .

Because of approximations in (16) and the inversion of the 3D Laplacian operator, the retrieved fields differ from the corresponding rhs forcing fields (not shown). Nonetheless, the perturbation (or ) in Figs. 9g–i shows some obvious consistencies with the rhs forcing fields in Figs. 9d–f: the strongest ascent coincides with the strongest forcing for ascent in (16), and weak descent coincides with the strongest forcing for descent. While the warm-frontal updraft strengthens with increased diabatic heating, it migrates progressively eastward from the northern flank (in T275-dry) to the northeastern edge (in T290-moist) of the warm sector. This increased phase difference between and at larger T0 tends to decrease term B in the moist cases.

Based on the above analysis, we propose a physical hypothesis for the reduced Ke and conversion efficiency with increasing T0 in the moist ISO simulations: diabatic heating along the warm-frontal cloud band locally enhances the low-level PV and , causing their corresponding maxima to strengthen and shift progressively eastward. While this effect enhances the magnitude of the dry-baroclinic forcing for vertical motion , it shifts the associated vertical motion eastward within the warm sector, which degrades the correlation between and . As a result, term B in (14) weakens at larger T0, which offsets the direct positive impacts of diabatic heating (term D) to leave the net C(Pe, Ke) largely unchanged. Thus, the moist cyclones are unable to intensify in warmer atmospheres despite their larger initial Pm.

A similar mechanism was cited in the aforementioned case study by Davis (1992). In that case, which also involved a type-B cyclone over land (the central United States), the diabatically forced eastward shift in the cyclone’s warm-frontal PV anomaly caused it to propagate faster than the corresponding upper-level PV anomaly. The resulting disruption of the phase-locking process ultimately led to cyclolysis.

4. Periodic waves

The ISO simulations represented one idealization of baroclinic life cycles—isolated cyclones developing within otherwise undisturbed flows. Another meaningful, and commonly studied, idealization is that of a periodic Rossby wave train, where repeated instances of the same wave can interact with each other (e.g., Whitaker and Davis 1994; BBP11; Booth et al. 2013). To evaluate the sensitivities of such waves to T0, and to compare their behavior to that of the ISO waves, we briefly analyze the PER simulations.

a. Overview

The baroclinic waves in the PER simulations differ subtly from those in the ISO simulations in that (i) the high-pressure regions flanking the cyclone are stronger, (ii) the cold front exhibits a noticeable forward tilt in the horizontal, and (iii) the cold front extends farther to the south (cf. Figs. 11 and 2). Both the dry and moist versions of the PER simulations attain 30%–50% larger Ke than the corresponding ISO simulations (cf. Figs. 12a,b and 3a,b). This generally increased Ke does not result from a deepening of the cyclone, as the magnitude of actually weakens relative to the ISO simulations (cf. Figs. 12e,f and 3e,f). Rather, it arises from the meridional elongation of the cyclones apparent in Fig. 11, which fills more of the analysis domain with wave energy. This elongation ultimately causes the cold front to reach the southern domain boundary late on day 7 (not shown), so we focus on results prior to that time. Moreover, in contrast to the cyclones in the ISO simulations, which deepen until day 7 then begin to dissipate, the cyclones in the PER simulations reach their minimum at around day 5 and decay very slowly thereafter.

Fig. 11.
Fig. 11.

As in Fig. 2, but for four corresponding PER simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Fig. 12.
Fig. 12.

As in Fig. 3, but for the PER simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The cyclones in the moist PER simulations again exhibit minimal sensitivity to T0 during their early growth phase, as the positive effects of increased diabatic heating at larger T0 offset the slower dry growth. They subsequently exhibit a stronger negative sensitivity to T0 than that in the corresponding moist ISO simulations, with the maximum Ke decreasing more prominently (~40%) over the range of T0. This negative sensitivity is even more pronounced in the conversion efficiency, which is largely insensitive to T0 in the dry cases but decreases strongly in the moist cases, by as much as 50% (cf. Figs. 12c,d and 3c,d). While the positive sensitivity of precipitation to T0 found in the moist ISO simulations persists in these simulations, the rate of increase of with T0 is greatly reduced (Table 2).

b. Energetics

The Ke budget of the PER simulations is computed identically to that of the ISO simulations, over a cyclone-following, limited-area zonal region (5400 km) that approximately matches the cyclone zonal wavelength. Most of the trends in this budget from the dry ISO simulations carry over to the dry PER simulations (Fig. 13), including a slight decrease in the Ke tendency during the cyclone growth phase at larger T0, stemming primarily from decreased C(Pe, Ke). Similarly, consistent with a slight weakening of the cyclone from T275-dry to T290-dry, both and also decrease in the latter case. Because the size of the analysis domain is nearly equal to the zonal wavelength of the cyclones, the net boundary fluxes are negligible and not shown.

Fig. 13.
Fig. 13.

As in Figs. 4a–d, but for the dry PER simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The pronounced decrease of Ke with increasing T0 in the moist PER simulations is caused primarily by a sharp increase in after day 5 (Fig. 14). This trend contrasts with all previous analyses, where decreased in magnitude at larger T0 owing to a weakening of the cyclone. Prior to day 6, C(Pe, Ke) behaves very similarly to that in the corresponding ISO simulations, with increased diabatic forcing and decreased baroclinic forcing roughly cancelling to give minimal net sensitivity (not shown). Subsequently, the weakening of the cyclone at larger T0, driven by , leads to a rapid decay of the other terms in the Ke budget.

Fig. 14.
Fig. 14.

As in Figs. 4a–d, but for the moist PER simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The increase of with T0 in the moist PER simulations is linked to the horizontal forward tilt of the cyclone (as seen in Figs. 11b and 11d), which causes a countergradient momentum flux that strengthens the mean zonal jet at the expense of the cyclone. The dominant term of in (12) is the product of Reynolds stresses with gradients in the zonal-mean zonal jet, or (Orlanski and Katzfey 1991). Because the meridional and vertical components of this term behave similarly (and the zonal term is zero), we focus on just the meridional term , the vertical integral of which over 400–800 hPa (the layer containing the most prominent differences) is shown at 0 h on day 7 in Fig. 15. This integral is generally more positive in T290-moist due to a larger correlation of , arising from the forward tilt of perturbation wind vectors and a stronger jet , with concomitantly stronger flanking gradients. These two factors are not independent: the latter arises from the former’s increase of .

Fig. 15.
Fig. 15.

Vertical integral over 400–800 hPa of at 0 h on day 7 of the (a) T275-moist and (b) T290-moist PER simulations (filled contours). Also shown are perturbation wind vectors and the mean zonal jet (thick lines, m s−1) over this layer, along with contours of surface pressure (interval: 10 hPa).

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

To explain the horizontal forward wave tilt in the moist PER simulations, we compare the evolution of upper-level (200–400 hPa) PV and difference wind vectors between the ISO and PER versions of the T290-moist case in Fig. 16. At 0 h on day 3, the two waves exhibit similar structures and diabatic heating distributions, with generally small differences in their wind fields (Figs. 16a,b). By the next day, an upper-level PV anomaly in the PER case, generated by diabatic heating along the warm front and carried downwind by the westerly jet, recirculates through the system to interact with the upper-level trough. Its anticyclonic flow is northerly along the upstream (western) flank of the trough, which tends to elongate the trough meridionally (Figs. 16c,d). Meanwhile, the zonally elongated warm-frontal diabatic PV anomaly at lower levels advects the surface thermal wave eastward (as seen in Figs. 9a–c). As a result of the differential propagation of the northern and southern ends of the wave, the wave develops a horizontal forward tilt over time, which increases with T0 as a result of increased diabatic PV generation.

Fig. 16.
Fig. 16.

Vertically averaged PV over 200–400 hPa (filled contours), along with 500–1000-hPa thickness (thin black lines, interval: 10 dam) and mean 100–1000-hPa diabatic heating rate (thick black line, contour 10−5 K s−1) for the (a),(c),(e) ISO and (b),(d),(f) PER versions of the T290-moist simulation. The difference mean 200–400-hPa wind vectors between the two simulations (PER minus ISO) are also shown in (b),(d),(f).

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

The meridional elongation of the upper-level trough explains the increased Ke in the PER simulations relative to the ISO simulations. However, in terms of surface pressure perturbations, the PER cyclones are relatively weak. This is because the phase-locking between lower- and upper-level PV anomalies in the ISO cyclones is not disrupted by the recirculating diabatic PV anomaly. Rather than being stretched meridionally, the upper trough wraps cyclonically around the low center, creating a downstream bulge in the trough that limits the horizontal wave tilt (Figs. 16e,f).

5. Discussion

a. Pe budget

While the foregoing analysis focused exclusively on the Ke budget, Ke constitutes just one of the four components (including Pm, Pe, Km, and Ke) of the Lorenz energy cycle (Lorenz 1955). Here we briefly inspect the Pe budget, which serves as the main pathway by which Pm is converted to Ke, and connects directly to the Ke budget through C(Pe, Ke). For ease of analysis, we derive an approximate form of the specific eddy potential energy pe
e17
where and are the initial specific volume and θ profiles, zonally averaged at . Multiplying the full thermodynamic energy equation by , subtracting the zonal-mean thermodynamic energy equation, and performing some manipulation gives
e18
with
e19
e20
e21
e22
where adv is the advection, is the diabatic generation, is the conversion from specific zonal available potential energy to , and .

As before, the integration operator in (11) is used to transform specific energy conversions in (18) into full energy conversions. The resulting Pe budget, shown for the T275-moist and T290-moist ISO simulations in Fig. 17, reveals clear parallels to the previous Ke analysis. The Pe tendencies in Fig. 17a are similar in magnitude to the corresponding Ke tendencies in Fig. 6a and again exhibit very weak dependence on T0 during the cyclone growth phase. This insensitivity reflects offsetting effects: a substantial increase in is balanced by a similar decrease in . Thus, the cancellation between competing diabatic and baroclinic processes as T0 is varied is consistent across both the Ke and Pe budgets. Similar cancellation is also found in the Pe budgets of the moist PER simulations (not shown).

Fig. 17.
Fig. 17.

As in Fig. 4, but for Pe budget terms in (18) in the moist ISO simulations.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

b. On the generalizability of the results

The above simulations correspond to a particular cyclogenesis configuration over land, where a large-amplitude upper-level PV anomaly propagates over a low-level baroclinic zone. This configuration, loosely based on the Petterssen type-B paradigm of cyclone development, is not representative of all midlatitude cyclones, which develop under widely different environmental conditions, with varying configurations of PV anomalies and low-level diabatic processes (e.g., Graf et al. 2017). Among the important environmental parameters neglected in these experiments is the basic-state barotropic horizontal wind shear, which can modify the strength of the warm front and its associated diabatic heating (e.g., Thorncroft et al. 1993). Because the feedbacks of diabatic heating on the parent cyclone depend on this and other parameters (e.g., Davis 1992; Davis et al. 1993), the broader applicability of our findings are uncertain.

While the increase in wave Ke due to moist processes in the ISO and PER simulations is consistent with many past studies (e.g., Balasubramanian and Yau 1994; Stoelinga 1996; Booth et al. 2013), the monotonic decrease in Ke with increasing T0 during the cyclone maturation phase is not. Although a nonmonotonic sensitivity to T0 has been found in GCMs (O’Gorman 2011; Pfahl et al. 2015), our simulations are more comparable to the NWP experiments of BBP11, where Ke monotonically increased with T0. BBP11’s setup differs from ours in terms of its (i) baroclinic jet, which was based on Polvani and Esler (2007) rather than Olson and Colle (2007); (ii) inclusion of weak meridional variation in initial RH, (iii) smaller domain size (Lx ≈ 4000 km); (iv) periodic initial perturbation, a small-amplitude thermal wave of wavelength ; and (v) an ocean surface with interactive heat fluxes. They also used a different NWP model (the Met Office Unified Model) with different physical parameterizations than those in our WRF experiments.

To reconcile the differences between our results and those of BBP11, we perform various experiments to bridge the gaps between the two experimental configurations. Starting with the moist ISO setup, we first replace the land surface with ocean, both with and without surface heat fluxes. For the latter, the sea surface temperature is set to the local surface air temperature minus 1 K. Compared to the reference case over land, Ke and are modified slightly, but their time evolution and sensitivity to T0 are largely unchanged (Figs. 18a,b). Next, we reduce to 4000 km and change the initial perturbation to a small-amplitude (1 K) sinusoidal thermal wave of wavelength . While these waves take longer to reach maturity, the negative sensitivity of wave intensity to T0 again holds (Figs. 18c,d).

Fig. 18.
Fig. 18.

Time series of Ke and (ΔSLP)min for various moist simulations, including (a),(b) ISO configuration with varied surface types; (c),(d) periodic waves in a small domain with an initial sinusoidal thermal perturbation (SIN); (e),(f) as in (c) and (d), but replacing the Olson and Colle (2007) jet with the Polvani and Esler (2007, or PE07) jet used in BBP11; and (g),(h) the ISO configuration with the Olson and Colle (2007) jet replaced with the PE07 jet. As shown in (a), blue and red lines correspond to T0 = 275 and 290 K, respectively, and solid, dashed, and dotted lines correspond to land, ocean, and ocean with interactive surface heat fluxes, respectively.

Citation: Journal of the Atmospheric Sciences 75, 1; 10.1175/JAS-D-17-0188.1

Keeping the small-domain configuration, we then replaced our initial jet with that of BBP11, with a similar maximum speed (45 m s−1) but a higher Z (nominally 12 km) and a stronger meridional gradient in both dry and moist static stability. Before day 6 of these simulations, both Ke and are 2–3 times larger for T0 = 290 K than for T0 = 275 K (Figs. 18e,f). However, the former begins to decay on day 6 while the latter continues to grow until day 8 and attain a larger peak intensity. This faster initial growth in the warmer case is consistent with BBP11, but the T0 sensitivity of wave-decay onset time is not. This discrepancy may relate to differences in the model representation of deep convection. For this configuration, CAPEmax exceeds 10 000 J kg−1 in the T0 = 290 K case, and intense parameterized convection develops by day 3. As in the PER simulations, the resulting large-amplitude anticyclonic PV anomalies aloft recirculate rapidly to upset the phase-locking process. For T0 = 275 K, CAPEmax = 0 and comparable diabatic PV anomalies require more time to develop.

As a final experiment, we return to the large-domain ISO setup but again replace our initial jet with that of BBP11. The isolated initial PV perturbation eliminates the negative feedbacks of recirculating PV anomalies, leading to a positive sensitivity of Ke and to T0 (Figs. 18g,h). This trend is the strongest for an interactive ocean surface, where a positive feedback between the cyclonic circulation, moist convection, and surface latent heat fluxes may preferentially energize the warmer wave (e.g., Emanuel 1986; Gray and Craig 1998). Although a detailed analysis of the mechanisms behind this trend is deferred to future work, we performed some additional tests to determine the sensitivity of this trend to the parameterizations of cloud diabatic heating, given the extreme CAPEmax in the T0 = 290-K case. Neither replacing the Kain–Fritsch cumulus scheme with the Tiedke scheme (which includes momentum adjustment), nor replacing the WSM6 microphysics scheme with the warm-rain Kessler scheme, changed the qualitative trends during the wave growth phase seen in Figs. 18g,h (not shown), suggesting a robustly positive T0 sensitivity for this initial configuration.

6. Conclusions

To evaluate the sensitivities of moist baroclinic waves to environmental temperature and moisture content, a series of idealized initial-value simulations of both moist and dry waves were conducted. These simulations used a configuration representative of Petterssen’s type-B cyclogenesis (Petterssen and Smebye 1971), where an isolated and strong upper-level PV anomaly propagates over a low-level baroclinic zone. With horizontal and vertical virtual-temperature gradients held fixed, the surface temperature at 45°N T0 was varied from 275 to 290 K. Relative to corresponding dry waves, the moist waves exhibited faster growth [based on eddy kinetic energy Ke and minimum surface pressure perturbation ]. However, the moist waves did not grow further as T0 was increased at constant relative humidity, despite a large attendant increase in atmospheric water content, moist available energy Pm, and CAPE.

The feedbacks of diabatic heating on moist wave intensity depended on the phase of the wave life cycle. During the early growth phase (days 2–3.5), increased cloud diabatic heating compensated for slightly reduced dry baroclinic instability in the warmer waves, allowing them to grow at a similar pace as the colder waves. However, during the maturation phase (days 3.5–7), the warmer waves grew more slowly than the colder waves, causing the peak Ke and maxima to decrease by ~15% over the simulated range of T0. As a result, the efficiency of conversion from initial Pm to wave Ke in these waves decreased rapidly with increasing T0.

The reduced conversion efficiency of the warmer moist waves was owing to an eastward shift in the low-level warm-frontal PV anomaly, driven by increased diabatic heating and moist convection along the frontal cloud band. It induced a corresponding eastward shift in the upper-level ridge, thereby increasing the downstream export of wave energy by ageostrophic geopotential fluxes. It also degraded the phasing between the buoyancy and the baroclinically forced vertical motion within the warm sector by shifting the latter progressively downstream of the former. The resulting indirect suppression of baroclinic conversion (from eddy potential energy to eddy kinetic energy) offset the direct enhancement of baroclinic conversion arising from diabatic heating.

Compared to isolated moist waves, periodic moist waves exhibited an even faster decrease in intensity and conversion efficiency at larger T0. This enhanced weakening was linked to the interaction of the upper-level trough with a recirculating anticyclonic upper-level diabatic PV anomaly originally generated along the warm front. As this PV anomaly caused the upper-level trough to elongate southward, the low-level diabatic cyclonic PV anomaly along the warm front advected the surface thermal wave eastward, causing the cyclone to develop a horizontal forward tilt. This tilt led to an increased conversion from wave Ke to mean kinetic energy, which strengthened the zonal jet at the expense of the cyclone. To our knowledge, such strong modifications of moist baroclinic wave growth by wave periodicity have not previously been documented and should be considered in future idealized modeling studies.

The finding that enhanced diabatic heating may feed back negatively on its parent cyclone is reminiscent of a case study by Davis (1992) over the central United States, where low-level cyclogenesis occurred in a similar type-B configuration to that used here. However, it differs from other case studies where diabatic heating led to major amplifications in cyclone development (e.g., Stoelinga 1996) and from previous initial-value experiments that found a monotonically positive sensitivity of wave Ke with T0 (Whitaker and Davis 1994; BBP11). Further experiments revealed that the simulated sensitivity can be substantially modified or even reversed by changing some combination of the initial jet structure, initial perturbation, and/or surface type. Thus, the sensitivities found herein likely only apply within a limited subset of parameter space. Given that no universal sensitivity of wave growth to T0 likely exists, experiments that sweep through the full parameter space, using a wide range of initial configurations of varying realism, are recommended to further examine this sensitivity.

Acknowledgments

The first author acknowledges funding from the Natural Sciences and Engineering Research Council (NSERC) Grant NSERC/RGPIN 418372-12. The numerical simulations were performed on the Guillimin supercomputer at McGill University, under the auspices of Calcul Québec and Compute Canada. We are grateful for insightful comments on this work from Paul O’Gorman [who also shared code for the efficient Pm calculation in Stansifer et al. (2017)], Bob Plant, Heini Wernli, John Methven, James Booth, and one anonymous reviewer.

REFERENCES

  • Adamson, D. S., S. E. Belcher, B. J. Hoskins, and R. S. Plant, 2006: Boundary-layer friction in midlatitude cyclones. Quart. J. Roy. Meteor. Soc., 132, 101124, https://doi.org/10.1256/qj.04.145.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ahmadi-Givi, F., G. C. Craig, and R. S. Plant, 2004: The dynamics of a midlatitude cyclone with very strong latent-heat release. Quart. J. Roy. Meteor. Soc., 130, 295323, https://doi.org/10.1256/qj.02.226.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Balasubramanian, G., and M. K. Yau, 1994: The effects of convection on a simulated marine cyclone. J. Atmos. Sci., 51, 23972417, https://doi.org/10.1175/1520-0469(1994)051<2397:TEOCOA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bengtsson, L., K. I. Hodges, and N. Keenlyside, 2009: Will extratropical storms intensify in a warmer climate? J. Climate, 22, 22762301, https://doi.org/10.1175/2008JCLI2678.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Booth, J. F., S. Wang, and L. Polvani, 2013: Midlatitude storms in a moister world: Lessons from idealized baroclinic life cycle experiments. Climate Dyn., 41, 787802, https://doi.org/10.1007/s00382-012-1472-3.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Boutle, I., S. E. Belcher, and R. S. Plant, 2011: Moisture transport in midlatitude cyclones. Quart. J. Roy. Meteor. Soc., 137, 360373, https://doi.org/10.1002/qj.783.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Büehler, D., and S. Pfahl, 2017: Potential vorticity diagnostics to quantify effects of latent heating in extratropical cyclones. Part I: Methodology. J. Atmos. Sci., 74, 35673590,https://doi.org/10.1175/JAS-D-17-0041.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Catto, J. L., L. C. Shaffrey, and K. I. Hodges, 2011: Northern Hemisphere extratropical cyclones in a warming climate in the HiGEM high-resolution climate model. J. Climate, 24, 53365352, https://doi.org/10.1175/2011JCLI4181.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Charney, J. G., 1947: The dynamics of long waves in a baroclinic westerly current. J. Meteor., 4, 136162, https://doi.org/10.1175/1520-0469(1947)004<0136:TDOLWI>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Colle, B. A., Z. Zhang, K. A. Lombardo, E. Chang, P. Liu, and M. Zhang, 2013: Historical evaluation and future prediction of eastern North American and western Atlantic extratropical cyclones in the CMIP5 models during the cool season. J. Climate, 26, 68826903, https://doi.org/10.1175/JCLI-D-12-00498.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., 1992: A potential-vorticity diagnosis of the importance of initial structure and condensational heating in observed extratropical cyclones. Mon. Wea. Rev., 120, 24092428, https://doi.org/10.1175/1520-0493(1992)120<2409:APVDOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., and K. A. Emanuel, 1991: Potential vorticity diagnostics of cyclogenesis. Mon. Wea. Rev., 119, 19291953, https://doi.org/10.1175/1520-0493(1991)119<1929:PVDOC>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., M. T. Stoelinga, and Y.-H. Kuo, 1993: The integrated effect of condensation in numerical simulations of extratropical cyclogenesis. Mon. Wea. Rev., 121, 23092330, https://doi.org/10.1175/1520-0493(1993)121<2309:TIEOCI>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • De Vries, H., J. Methven, T. A. Frame, and B. J. Hoskins, 2010: Baroclinic waves with parameterized effects of moisture interpreted using Rossby wave components. J. Atmos. Sci., 67, 27662784, https://doi.org/10.1175/2010JAS3410.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1, 3352, https://doi.org/10.3402/tellusa.v1i3.8507.

  • Emanuel, K. A., 1986: An air-sea interaction theory for tropical cyclones. Part I: Steady-state maintenance. J. Atmos. Sci., 43, 585605, https://doi.org/10.1175/1520-0469(1986)043<0585:AASITF>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K. A., M. Fantini, and A. J. Thorpe, 1987: Baroclinic instability in an environment of small stability to slantwise moist convection. Part I: Two-dimensional models. J. Atmos. Sci., 44, 15591573, https://doi.org/10.1175/1520-0469(1987)044<1559:BIIAEO>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Fantini, M., 2004: Baroclinic instability of a zero-PVE jet: Enhanced effects of moisture on the life cycle of midlatitude cyclones. J. Atmos. Sci., 61, 12961307, https://doi.org/10.1175/1520-0469(2004)061<1296:BIOAZJ>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Farrell, B., 1985: Transient growth of damped baroclinic waves. J. Atmos. Sci., 42, 27182727, https://doi.org/10.1175/1520-0469(1985)042<2718:TGODBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Farrell, B., 1989: Optimal excitation of baroclinic waves. J. Atmos. Sci., 46, 11931206, https://doi.org/10.1175/1520-0469(1989)046<1193:OEOBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Govindasamy, B., and S. T. Garner, 1997: The equilibration of short baroclinic waves. J. Atmos. Sci., 54, 28502871, https://doi.org/10.1175/1520-0469(1997)054<2850:TEOSBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Graf, M. A., H. Wernli, and M. Sprenger, 2017: Objective classification of extratropical cyclogenesis. Quart. J. Roy. Meteor. Soc., 143, 10471061, https://doi.org/10.1002/qj.2989.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gray, S. L., and G. C. Craig, 1998: A simple theoretical model for the intensification of tropical cyclones and polar lows. Quart. J. Roy. Meteor. Soc., 124, 919947, https://doi.org/10.1002/qj.49712454713.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Holton, J. R., 1992: An Introduction to Dynamic Meteorology. 3rd ed. Academic Press, 511 pp.

  • Hong, S.-Y., and J.-O. J. Lim, 2006: The WRF single-moment 6-class microphysics scheme (WSM6). J. Korean Meteor. Soc., 42, 129151.

  • Hong, S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 23182341, https://doi.org/10.1175/MWR3199.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hoskins, B. J., I. Draghici, and H. C. Davies, 1978: A new look at the ω-equation. Quart. J. Roy. Meteor. Soc., 104, 3138, https://doi.org/10.1002/qj.49710443903.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • 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, 877946, https://doi.org/10.1002/qj.49711147002.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • IPCC, 2013: Climate Change 2013: The Physical Science Basis. Cambridge University Press, 1535 pp., https://doi.org/10.1017/CBO9781107415324.

    • Crossref
    • Export Citation
  • Kain, J. S., 2004: The Kain–Fritsch convective parameterization: An update. J. Appl. Meteor., 43, 170181, https://doi.org/10.1175/1520-0450(2004)043<0170:TKCPAU>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lackmann, G., 2011: Midlatitude Synoptic Meteorology: Dynamics, Analysis and Forecasting. Amer. Meteor. Soc., 345 pp.

    • Crossref
    • Export Citation
  • Lapeyre, G., and I. M. Held, 2004: The role of moisture in the dynamics and energetics of turbulent baroclinic eddies. J. Atmos. Sci., 61, 16931710, https://doi.org/10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 1955: Available potential energy and the maintenance of the general circulation. Tellus, 7, 157167, https://doi.org/10.3402/tellusa.v7i2.8796.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mak, M., 1994: Cyclogenesis in a conditionally unstable moist baroclinic atmosphere. Tellus, 46A, 1433, https://doi.org/10.3402/tellusa.v46i1.15424.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Marciano, C. G., G. M. Lackmann, and W. A. Robinson, 2015: Changes in U.S. East Coast cyclone dynamics with climate change. J. Climate, 28, 468484, https://doi.org/10.1175/JCLI-D-14-00418.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merlis, T. M., and T. Schneider, 2009: Scales of linear baroclinic instability and macroturbulence in dry atmospheres. J. Atmos. Sci., 66, 18211833, https://doi.org/10.1175/2008JAS2884.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • O’Gorman, P. A., 2011: The effective static stability experienced by eddies in a moist atmosphere. J. Atmos. Sci., 68, 7590, https://doi.org/10.1175/2010JAS3537.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • O’Gorman, P. A., T. M. Merlis, and M. S. Singh, 2017: Increase in the skewness of extratropical vertical velocities with climate warming: Fully nonlinear simulations versus moist baroclinic instability. Quart. J. Roy. Meteor. Soc., https://doi.org/10.1002/qj.3195, in press.

    • Search Google Scholar
    • Export Citation
  • Olson, J. B., and B. A. Colle, 2007: A modified approach to initialize an idealized extratropical cyclone within a mesoscale model. Mon. Wea. Rev., 135, 16141624, https://doi.org/10.1175/MWR3364.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Orlanski, I., and J. Katzfey, 1991: The life cycle of a cyclone wave in the Southern Hemisphere. Part I: Eddy energy budget. J. Atmos. Sci., 48, 19721998, https://doi.org/10.1175/1520-0469(1991)048<1972:TLCOAC>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Petterssen, S., and S. J. Smebye, 1971: On the development of extratropical cyclones. Quart. J. Roy. Meteor. Soc., 97, 457482, https://doi.org/10.1002/qj.49709741407.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pfahl, S., P. A. O’Gorman, and M. S. Singh, 2015: Extratropical cyclones in idealized simulations of changed climates. J. Climate, 28, 93739392, https://doi.org/10.1175/JCLI-D-14-00816.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Polvani, L. M., and J. G. Esler, 2007: Transport and mixing of chemical air masses in idealized baroclinic life cycles. J. Geophys. Res., 112, D23102, https://doi.org/10.1029/2007JD008555.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Reed, R. J., A. J. Simmons, M. D. Albright, and P. Unden, 1988: The role of latent heat release in explosive cyclogenesis: Three examples based on ECMWF operational forecasts. Wea. Forecasting, 3, 217229, https://doi.org/10.1175/1520-0434(1988)003<0217:TROLHR>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Schär, C., and H. Wernli, 1993: Structure and evolution of an isolated semi-geostrophic cyclone. Quart. J. Roy. Meteor. Soc., 119, 5790, https://doi.org/10.1002/qj.49711950904.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Schär, C., C. Frei, D. Lüthi, and H. C. Davies, 1996: Surrogate climate-change scenarios for regional climate models. Geophys. Res. Lett., 23, 669672, https://doi.org/10.1029/96GL00265.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Schemm, S., H. Wernli, and L. Papritz, 2013: Warm conveyor belts in idealized moist baroclinic wave simulations. J. Atmos. Sci., 70, 627652, https://doi.org/10.1175/JAS-D-12-0147.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Schneider, T., P. A. O’Gorman, and X. J. Levine, 2010: Water vapor and the dynamics of climate changes. Rev. Geophys., 48, RG3001, https://doi.org/10.1029/2009RG000302.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Simmons, A. J., and B. J. Hoskins, 1976: Baroclinic instability on the sphere: Normal models of the primitive and quasi-geostrophic equations. J. Atmos. Sci., 33, 14541477, https://doi.org/10.1175/1520-0469(1976)033<1454:BIOTSN>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Skamarock, W. C., and Coauthors, 2008: A description of the Advanced Research WRF version 3. NCAR Tech. Note NCAR/TN-475+STR, 113 pp., https://dx.doi.org/10.5065/D68S4MVH.

    • Crossref
    • Export Citation
  • Stansifer, E. M., P. A. O’Gorman, and J. I. Holt, 2017: Accurate computation of moist available potential energy with the Munkres algorithm. Quart. J. Roy. Meteor. Soc., 143, 288292, https://doi.org/10.1002/qj.2921.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Stoelinga, M. T., 1996: A potential vorticity-based study of the role of diabatic heating and friction in a numerically simulated baroclinic cyclone. Mon. Wea. Rev., 124, 849874, https://doi.org/10.1175/1520-0493(1996)124<0849:APVBSO>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Takayabu, I., 1991: “Coupling development”: An efficient mechanism of the development of extratropical cyclones. J. Meteor. Soc. Japan, 69, 609628, https://doi.org/10.2151/jmsj1965.69.6_609.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Thorncroft, C. D., B. J. Hoskins, and M. E. McIntyre, 1993: Two paradigms of baroclinic-wave life-cycle behaviour. Quart. J. Roy. Meteor. Soc., 119, 1755, https://doi.org/10.1002/qj.49711950903.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Trenberth, K. E., 1978: On the interpretation of the diagnostic quasi-geostrophic omega equation. Mon. Wea. Rev., 106, 131137, https://doi.org/10.1175/1520-0493(1978)106<0131:OTIOTD>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and C. A. Davis, 1994: Cyclogenesis in a saturated environment. J. Atmos. Sci., 51, 889907, https://doi.org/10.1175/1520-0469(1994)051<0889:CIASE>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Willison, J., W. A. Robinson, and G. M. Lackmann, 2013: The importance of resolving mesoscale latent heating in the North Atlantic storm track. J. Atmos. Sci., 70, 22342250, https://doi.org/10.1175/JAS-D-12-0226.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zappa, G., L. C. Shaffrey, K. I. Hodges, P. G. Sansom, and D. B. Stephenson, 2013: A multimodel assessment of future projections of north Atlantic and European extratropical cyclones in the CMIP5 climate models. J. Climate, 26, 58465862, https://doi.org/10.1175/JCLI-D-12-00573.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
Save
  • Adamson, D. S., S. E. Belcher, B. J. Hoskins, and R. S. Plant, 2006: Boundary-layer friction in midlatitude cyclones. Quart. J. Roy. Meteor. Soc., 132, 101124, https://doi.org/10.1256/qj.04.145.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ahmadi-Givi, F., G. C. Craig, and R. S. Plant, 2004: The dynamics of a midlatitude cyclone with very strong latent-heat release. Quart. J. Roy. Meteor. Soc., 130, 295323, https://doi.org/10.1256/qj.02.226.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Balasubramanian, G., and M. K. Yau, 1994: The effects of convection on a simulated marine cyclone. J. Atmos. Sci., 51, 23972417, https://doi.org/10.1175/1520-0469(1994)051<2397:TEOCOA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bengtsson, L., K. I. Hodges, and N. Keenlyside, 2009: Will extratropical storms intensify in a warmer climate? J. Climate, 22, 22762301, https://doi.org/10.1175/2008JCLI2678.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Booth, J. F., S. Wang, and L. Polvani, 2013: Midlatitude storms in a moister world: Lessons from idealized baroclinic life cycle experiments. Climate Dyn., 41, 787802, https://doi.org/10.1007/s00382-012-1472-3.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Boutle, I., S. E. Belcher, and R. S. Plant, 2011: Moisture transport in midlatitude cyclones. Quart. J. Roy. Meteor. Soc., 137, 360373, https://doi.org/10.1002/qj.783.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Büehler, D., and S. Pfahl, 2017: Potential vorticity diagnostics to quantify effects of latent heating in extratropical cyclones. Part I: Methodology. J. Atmos. Sci., 74, 35673590,https://doi.org/10.1175/JAS-D-17-0041.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Catto, J. L., L. C. Shaffrey, and K. I. Hodges, 2011: Northern Hemisphere extratropical cyclones in a warming climate in the HiGEM high-resolution climate model. J. Climate, 24, 53365352, https://doi.org/10.1175/2011JCLI4181.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Charney, J. G., 1947: The dynamics of long waves in a baroclinic westerly current. J. Meteor., 4, 136162, https://doi.org/10.1175/1520-0469(1947)004<0136:TDOLWI>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Colle, B. A., Z. Zhang, K. A. Lombardo, E. Chang, P. Liu, and M. Zhang, 2013: Historical evaluation and future prediction of eastern North American and western Atlantic extratropical cyclones in the CMIP5 models during the cool season. J. Climate, 26, 68826903, https://doi.org/10.1175/JCLI-D-12-00498.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., 1992: A potential-vorticity diagnosis of the importance of initial structure and condensational heating in observed extratropical cyclones. Mon. Wea. Rev., 120, 24092428, https://doi.org/10.1175/1520-0493(1992)120<2409:APVDOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., and K. A. Emanuel, 1991: Potential vorticity diagnostics of cyclogenesis. Mon. Wea. Rev., 119, 19291953, https://doi.org/10.1175/1520-0493(1991)119<1929:PVDOC>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Davis, C. A., M. T. Stoelinga, and Y.-H. Kuo, 1993: The integrated effect of condensation in numerical simulations of extratropical cyclogenesis. Mon. Wea. Rev., 121, 23092330, https://doi.org/10.1175/1520-0493(1993)121<2309:TIEOCI>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • De Vries, H., J. Methven, T. A. Frame, and B. J. Hoskins, 2010: Baroclinic waves with parameterized effects of moisture interpreted using Rossby wave components. J. Atmos. Sci., 67, 27662784, https://doi.org/10.1175/2010JAS3410.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1, 3352, https://doi.org/10.3402/tellusa.v1i3.8507.

  • Emanuel, K. A., 1986: An air-sea interaction theory for tropical cyclones. Part I: Steady-state maintenance. J. Atmos. Sci., 43, 585605, https://doi.org/10.1175/1520-0469(1986)043<0585:AASITF>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K. A., M. Fantini, and A. J. Thorpe, 1987: Baroclinic instability in an environment of small stability to slantwise moist convection. Part I: Two-dimensional models. J. Atmos. Sci., 44, 15591573, https://doi.org/10.1175/1520-0469(1987)044<1559:BIIAEO>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Fantini, M., 2004: Baroclinic instability of a zero-PVE jet: Enhanced effects of moisture on the life cycle of midlatitude cyclones. J. Atmos. Sci., 61, 12961307, https://doi.org/10.1175/1520-0469(2004)061<1296:BIOAZJ>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Farrell, B., 1985: Transient growth of damped baroclinic waves. J. Atmos. Sci., 42, 27182727, https://doi.org/10.1175/1520-0469(1985)042<2718:TGODBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Farrell, B., 1989: Optimal excitation of baroclinic waves. J. Atmos. Sci., 46, 11931206, https://doi.org/10.1175/1520-0469(1989)046<1193:OEOBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Govindasamy, B., and S. T. Garner, 1997: The equilibration of short baroclinic waves. J. Atmos. Sci., 54, 28502871, https://doi.org/10.1175/1520-0469(1997)054<2850:TEOSBW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Graf, M. A., H. Wernli, and M. Sprenger, 2017: Objective classification of extratropical cyclogenesis. Quart. J. Roy. Meteor. Soc., 143, 10471061, https://doi.org/10.1002/qj.2989.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gray, S. L., and G. C. Craig, 1998: A simple theoretical model for the intensification of tropical cyclones and polar lows. Quart. J. Roy. Meteor. Soc., 124, 919947, https://doi.org/10.1002/qj.49712454713.