1. Introduction
African easterly waves (AEWs) dominate the synoptic-scale circulation over North Africa during summer. AEWs occur every 2–6 days and travel westward along two tracks that straddle the African easterly jet (AEJ) (Pytharoulis and Thorncroft 1999), which is located at midtroposphere and aligned along ~15°N. The AEJ is the major source of energy for the AEWs and therefore affects their spatial–temporal evolution (Burpee 1972, 1974; Reed et al. 1977, 1988).
The AEWs that migrate along the north track extend their influence into the Sahara Desert, where they contribute to the emission, mixing, and transport of Saharan mineral dust aerosols (Jones et al. 2003; Knippertz and Todd 2010; Grogan and Thorncroft 2019; Nathan et al. 2019). The dust episodically coalesces into synoptic-scale plumes that absorb, scatter, and emit radiation, processes that affect the strength and structure of both the AEJ (Tompkins et al. 2005; Chen et al. 2010; Reale et al. 2011) and the AEWs (see Table 1).
Select papers on dust-coupled AEWs.
As discussed in Bercos-Hickey et al. (2017), dust-induced changes to the AEWs can be envisaged as operating along two pathways. One pathway is associated with the zonal-mean portion of the dust field (pathway I) and the other pathway with the eddy portion (pathway II).1 Along pathway I the zonal-mean dust alters the zonal-mean temperature field, which, due to thermal wind balance, alters the location, strength, and structure of the AEJ. As shown by Grogan et al. (2017), the AEWs then experience the effects of the zonal-mean dust through their interaction with the dust-modified AEJ.
Along pathway II, as shown by Nathan et al. (2017), the dust-induced eddy diabatic heating rate is a function of the zonal-mean aerosol optical depth (AOD), and the advection of zonal-mean dust by the eddy wind field. Together these radiative–dynamical processes change the growth, propagation, and structure of the AEWs (Grogan et al. 2016; Bercos-Hickey et al. 2017; Nathan et al. 2017).
Pathways I and II together constitute the dust-modified wave–mean flow interaction, a fundamentally nonlinear process (Grogan et al. 2017). If the background distributions of wind, temperature, and dust are specified, as in a linear model, then only pathway II operates on the dust-modified AEWs.
Grogan et al. (2016), for example, examined pathway II. To do so, they coupled an idealized version of the Weather Research and Forecasting (WRF) Model with a simplified dust model to understand the radiative–dynamical feedbacks that control the dust-modified linear dynamics of AEWs. The dust model, which was coupled radiatively to the WRF Model, included the effects of advection and sedimentation. The WRF-Dust model was used to conduct linear simulations around a zonal-mean background state that was chosen consistent with the observed zonal-mean distributions of wind, temperature, and dust over North Africa during summer. The model AEJ was chosen supercritical with respect to the threshold for combined barotropic–baroclinic instability. Grogan et al. (2016) showed that for the fastest-growing AEW, the growth rate increased monotonically with AOD, ranging from 15% to 90% for AODs ranging from 1.0 to 2.5. To explain the model results, Grogan et al. (2016) derived an analytical expression for the generation of eddy available potential energy (APE) that showed how the dust most effectively feeds back on the AEWs to affect their growth. The expression shows that the maximum generation of eddy APE occurs where the Doppler-shifted frequency2 is minimized and the meridional gradient of zonal-mean dust is maximized. This unique region occurs between the AEJ axis (~15°N) and the maximum in the dust distribution (~20°N).
Nathan et al. (2017) combined a theoretical framework with WRF-Dust model experiments to examine the subcritical destabilization of AEWs by the dust. The theoretical framework yielded analytical expressions for the dust-modified growth rate and frequency of the modeled AEWs. The expressions not only expose how the dust couples to the circulation to affect the AEWs, but they also serve as a tool for predicting what background wind and dust distributions can destabilize the AEWs. The theoretical predictions were confirmed through WRF-Dust model simulations, which showed that otherwise neutral AEWs can be destabilized by the dust, with growth rates similar to those obtained in supercritical AEJs (cf. Rennick 1976; Thorncroft and Hoskins 1994; Grogan et al. 2016).
The idealized studies of Grogan et al. (2016) and Nathan et al. (2017) examined different background jets—one supercritical and the other subcritical—yet they shared two key results: the dust enhances the growth of AEW-like waves; and the maximum generation of eddy APE is located between the AEJ axis and the dust maximum. Despite the simplicity of their models, the results are robust. This was confirmed by Bercos-Hickey et al. (2017), who used the full nonlinear, three-dimensional version of the WRF-Dust model, which includes processes excluded in the linear studies cited above, that is, wind-generated emission of dust, cloud microphysical processes, subgrid cumulus and boundary layer mixing of dust, and wet and dry deposition. Bercos-Hickey et al. (2017) showed that for an AEJ representative of summer 2006, the dust strengthens the AEWs and maximizes both the generation of eddy APE and the baroclinic energy conversion between the AEJ and dust maximum. The fact that the full WRF-Dust model reproduces the key results from the idealized version lends confidence in the ability of the idealized WRF-Dust model to not only expose the dust-modified dynamics of the AEWs, but to also replicate the results obtained from more complicated models.
The above dust–AEW studies each conducted their analyses around a single, though observationally representative, AEJ. Observations show, however, that the AEJ has pronounced intraseasonal variability (Newell and Kidson 1984; Afiesimama 2007) as well as interannual variability (Grist and Nicholson 2001; Grist et al. 2002; Dezfuli and Nicholson 2011). Such variability manifests in structural changes to the AEJ, which has been shown in previous idealized, dust-free modeling studies to significantly affect the linear dynamics of AEWs (Rennick 1981; Kwon 1989; Thorncroft 1995; Paradis et al. 1995).
It is unclear, however, to what extent structural changes in the AEJ mediate the effects of dust on the growth of AEWs. For example, are the dust-modified growth rates of the AEWs more sensitive to changes in the vertical or meridional shear of the background jet? How does the jet structure affect the dust modified energetics of the AEWs? And is there an optimal jet structure that maximizes the effects of the dust on the growth of AEWs?
To answer these questions, we combine an analytical analysis of the dust-modified energetics with idealized WRF-Dust model experiments. Such an approach aids in the exposure of important physical–dynamical processes, processes that may be obscured in observational data or difficult to identify in complex models. Moreover, the idealized WRF-Dust model is easily modified and thus is particularly useful for sensitivity analyses of different background jets, such as those examined in this study.
2. Model, basic states, and experimental design
a. Model
We use the same model as Grogan et al. (2016). The model has a WRF dynamical core (v3.7) that is coupled to the dust model developed by Chen et al. (2010, 2015). The dust model simulates the advection and sedimentation of 12-bin dust mixing ratios; the particles’ radii range from 0.15 to 5.0 μm. The radiative properties of the dust are calculated using the NASA Goddard Space Flight Center radiation model (Chou and Suarez 1999; Chou et al. 2001), which uses the Optical Properties of Aerosols and Clouds (OPAC) software package (Hess et al. 1998) to obtain the dust optical properties—extinction, single scattering albedo, and asymmetric parameter. The radiation model computes the daily averaged dust-induced heating rates, which includes the absorption and scattering due to dust, and the reabsorption of dust emission by other constituents.
The domain is a global channel projected on an equidistant cylindrical grid. In the horizontal, the resolution is 0.5° and the boundary conditions are periodic in the zonal direction and symmetric at the channel sidewalls (40°N and 10°S). In the vertical, there are 50 levels that extend from a flat surface up to 100 hPa, with no-slip conditions at the top and bottom. To reduce noise and boundary reflections, we impose a 30-min hyperdiffusion on the wind fields throughout the model domain, and a Rayleigh damping on the wind and temperature fields at the model top (300–100 hPa).
b. Basic states
Our linear stability analysis pivots on a zonal-mean background jet that is identical to that used in Grogan et al. (2016) (Fig. 1). The jet is consistent with the observed AEJ: it is symmetric in latitude, asymmetric in height, and has a maximum easterly wind of 15 m s−1 centered at 15°N and ~650 hPa. The linear stability of this model AEJ will serve as our reference, about which we construct four other background states: two correspond to either a 25% reduction or a 25% enhancement in the maximum meridional shear zones, which are located north and south of the AEJ axis. The other two AEJs correspond to either a 25% reduction or a 25% enhancement of the maximum vertical shear below 650 hPa. These are the percentages used by Rennick (1981) and Paradis et al. (1995) in their idealized modeling studies of the sensitivity of the stability of the AEJ to changes in its structure.
We denote the five background jets by UREF (reference jet),
Maximum values of zonal-mean wind speed, meridional shear, and vertical shear for each of the five basic states. The maximum meridional shear on the north and south flanks are the same in magnitude but opposite in sign.
Figure 2 shows the background distribution of the dust mixing ratio and daily averaged dust-induced heating rate. As in Grogan et al. (2016), each of the 12 particle sizes that constitute the dust field have the same horizontal distribution, which is symmetric in the horizontal and centered at 20°N. In contrast to Grogan et al. (2016), we use a more realistic vertical profile and particle size distribution for the dust field. The vertical profile is guided by dust extinction profiles from CALIPSO lidar measurements over North Africa from Yu et al. (2010) (see their Fig. 10), while the particles follow the more realistic logarithmic size distribution described in Kok (2011, see their Fig. 2). The size distributions are scaled to produce a maximum AOD of ~1.0, which yields a dust maximum mixing ratio of ~800 μg kg−1 at 20°N near the surface (Fig. 2a).
Figure 2b shows the basic-state daily averaged dust-induced heating rate. The weak dust-induced warming at the surface obtained by Grogan et al. (2016) is absent in Fig. 2b, a consequence of our choosing a more realistic logarithmic particle size distribution. More importantly, the absence of surface heating is more consistent with other dust-induced heating rate profiles over North Africa (e.g., Quijano et al. 2000). Above the surface, the heating rate increases with height, reaching a maximum at ~700 hPa, which is ~50 hPa lower and ~15% stronger than the maximum heating rate given in Grogan et al. (2016). Above 700 hPa, the heating rates decrease with height.
c. Experimental design
For each AEJ presented in Table 2, numerical experiments are run without dust (NODUST) and with dust (DUST). For the NODUST experiments, the radiative effects of the dust are decoupled from the thermodynamics so that the dust is simply a passive tracer. For the DUST experiments, the dust is interactive; that is, it absorbs, scatters, and emits radiation to affect the thermal field, which, in turn, affects the circulation and the transport of dust. Because all of the other parameterizations, such as microphysics, convection, and boundary layer physics, are deactivated in the model, the differences between NODUST and DUST isolate the direct dust-radiative feedbacks on the linear dynamics of the AEWs.
The linear stability characteristics of the AEWs are obtained using the same methodology as in previous studies (Thorncroft and Hoskins 1994; Simmons and Hoskins 1976; Grogan et al. 2016). For each experiment, a wave packet is superimposed onto the zonal-mean wind field. The packet contains zonal wavelengths spanning 2500–5000 km (nondimensional wavenumbers 8–16), which are typical for observed AEWs (Burpee 1975). The initial amplitude of the wave packet is small (~10−4 m s−1), which ensures linearity throughout the simulation [see Grogan et al. (2016) for details]. The model is integrated in time until the domain-averaged perturbation energy for each wavelength grows exponentially to an accuracy of 10−3 for at least 12 h. A fast Fourier transform algorithm is used to obtain the desired disturbance structures from each field. The fields are then scaled to produce a peak meridional wind of 5 m s−1.
3. Analytical analysis
List of variables.
Except for slight differences in notation, (3.3a)–(3.3f) are the same as in Hsieh and Cook (2007). But in contrast to their study, our global energy conversions are devoid of boundary terms, a consequence of our boundary conditions: zonal periodicity, and the application of the kinematic boundary condition at the channel sidewalls and at the lower and upper boundaries.
If we note that A ≫ B in (3.14), which is confirmed by our numerical calculations shown in section 4, we can then write
4. Numerical results
a. Growth rates and phase speeds
Figure 3 shows the growth rates as a function of zonal wavenumber without dust (left column) and with dust (right column) for each of the background AEJs. Figure 3a shows that for UREF, the NODUST growth rates are, as expected, the same as those in Grogan et al. (2016); for the fastest-growing AEW (k = 12) the growth rate is 0.34 day−1. Figure 3a also shows that adjustments to the peak meridional shear affects the growth rates across the wave spectrum, and shifts the zonal scale of the fastest-growing wave. For example, compared to UREF, the growth rates obtained from
Figure 3b compares the NODUST growth rates obtained from
Comparison of Figs. 3a and 3b shows that adjustments to the meridional shear produce greater responses to the growth rates than adjustments to the vertical shear. This result is consistent with Rennick (1981) but not with Paradis et al. (1995), who attributed the difference between their study and Rennick’s (1981) to structural differences between their background jets. Here, however, we find that the growth rates are less sensitive to changes in the vertical shear, as in Rennick (1981), yet our reference jet is closer to that of Paradis et al. (1995), who found that the growth rates were more sensitive to adjustments in the vertical shear. To resolve this apparent conundrum, we compared our model with that of Paradis et al. (1995) and found that although our jets are similar, our mean vertical temperature profiles are not; our vertical temperature profile produces a more statically stable lower troposphere. To examine how the differences in the static stability between Paradis et al. (1995) and our study affect the response of the growth rates to adjustments in the vertical shear, we carried out sensitivity tests that involved reducing the static stability for the NODUST simulations. The reduced static stability increased the baroclinic energy conversions, which resulted in growth rates that were more sensitive to adjustments in the vertical shear, consistent with Paradis et al. (1995).
Figures 3c and 3d compare the DUST growth rates for UREF with the perturbed meridional shear jets
Growth rates, easterly phase speeds, and zonal wavelengths of the fastest-growing AEW for each of the AEJs listed in Table 2.
The phase speeds for the NODUST and DUST were also calculated for each AEJ profile. As in previous dust-free studies (e.g., Thorncroft 1995), the phase speeds are fastest for flows with reduced meridional and vertical shear (i.e., reduced
b. Energetics
The global and local energetics are calculated for zonal wave k = 12 (wavelength = 3300 km). The energetics of this wave is representative of the other zonal waves examined in the previous subsection.
Table 5 shows the global energy conversions for the NODUST and DUST experiments. There are three general results. First, for each AEJ, with or without dust, the barotropic conversions are larger than the baroclinic conversions, that is, CE/CK < 1, which agrees with Grogan et al. (2016) and the studies cited in their Table 2. For this study, CE/CK is largest for
Global energy conversions (1 × 10−3 W m−2) for the k = 12 NODUST and DUST experiments. For NODUST, GE = 0.0 for each AEJ.
Figures 4–7 show the local energetics,
Figure 4 shows
Figures 5 and 6 show, respectively,
Figure 7 shows
Figure 7 also shows that the analytical analysis predicts the location of peak
5. Conclusions
Analytical and numerical approaches are used to examine how structural changes in the African easterly jet (AEJ) mediate the direct radiative effects of Saharan mineral dust aerosols on the growth and propagation of African easterly waves (AEWs). The analytical approach hinged on deriving an expression for the generation of eddy available potential energy (APE), an expression that exposes how the structure of the AEJ, dust distribution and Doppler-shifted frequency combine to affect the energetics of the AEWs. The analytical expression was used to interpret the numerical results, which were obtained by coupling a simplified version of the Weather Research and Forecasting (WRF) Model with a simplified conservation equation for the eddy dust field. The WRF-Dust model was used to conduct linear simulations around a zonal-mean background state that was chosen consistent with observations over North Africa during summer. The peak meridional and vertical shears of the background AEJ were varied and the linear dynamics of the AEWs subsequently examined.
The key findings are as follows: (i) Irrespective of the AEJ structure or the zonal scale of the AEWs, dust increases the growth rates of the AEWs. (ii) The growth rates of the AEWs are optimized when the ratio of baroclinic to barotropic energy conversions is largest. (iii) When the energy conversions are sufficiently large, the zonal scale of the fastest-growing AEW shortens.
For each of the AEJs that were examined, the dust generates eddy APE (GE) at midlevels north of the AEJ axis (850–650 hPa, 15°–20°N) and increases the conversion of zonal to eddy APE (CA) and baroclinic energy conversion (CE) in the same region. Grogan and Thorncroft (2019) used MERRA-2 reanalysis to infer the distribution of GE and CE associated with dust-coupled AEWs. Our model calculations qualitatively agree with their results.
For AEJs with enhanced vertical shear and reduced meridional shear, the fastest-growing AEWs have the largest increase in growth rate (28%–35%), a consequence of the larger increase in CA, CE, and barotropic energy conversions (CK) For AEJs with reduced vertical shear and enhanced meridional shear, however, the fastest-growing waves have smaller increases (12%–18%), a consequence of smaller increases in CA and CE, and decreases in CK.
As discussed in section 3 and noted above, the interpretation of the energetics results obtained from the WRF-Dust model was aided by our analytical analysis of GE. The analysis shows that the strength of GE depends on the zonal-mean transmissivity, and the zonal-mean meridional and vertical gradients of the background dust field, which are modulated by the Doppler-shifted frequency. The expression for GE can be interpreted as a correction to CA, where the gradients of zonal-mean dust modify the preexisting zonal-mean temperature gradients. For the background states examined in this study, CA + GE shows that the dust always increases the production of eddy APE (and thus growth rates). The increase is due to the positive meridional dust gradient north of the AEJ axis, which aligns with the low-level positive meridional temperature gradient, which is also in the vicinity of a critical surface. Consequently, the differential heating associated with the dust gradient reinforces the temperature gradient, which is also where the equatorward heat fluxes of the unstable AEWs are largest (Norquist et al. 1977). Therefore, knowledge of the background distributions of wind, temperature and dust over Africa can be used as a predictive tool to assess the influence of dust radiative effects on the growth of AEWs.
Here we have examined the linear response of AEWs to changes in the structure of AEJs that are supercritical with respect to combined barotropic–baroclinic instability. A logical extension of this work would be to examine the linear dynamics of AEWs in subcritical background flows, which, as shown by Nathan et al. (2017), can support the dust-induced growth of otherwise neutral AEWs. Nathan et al. (2017), however, only considered a single, though observationally representative, AEJ. It would therefore be interesting to determine which subcritical AEJ structures optimize the dust-induced instabilities of the AEWs. Equally important is the consideration of the dust-modified nonlinear evolution of AEWs in different initial AEJ configurations. In such a problem the background distributions of zonal-mean and eddy dust would each feed back on the circulation to affect the life cycles of the AEWs. In Grogan et al. (2017), for example, it was shown for an initial supercritical jet, the total dust field (mean plus eddy) enhanced the linear growth, weakened the nonlinear stabilization, and caused larger peak amplitude. Throughout most of the AEW life cycle, the Eliassen–Palm flux divergence and residual mean circulation were both enhanced, but operated in opposition to each other. To what extent these results are sensitive to changes in the initial structure of the AEJ is unclear, but will be examined in a future study.
Acknowledgments
The authors thank Emily Bercos-Hickey and Scott Nelson for their comments on the manuscript. We also acknowledge high-performance computing support from Cheyenne (doi:10.5065/D6RX99HX) provided by NCAR’s Computational and Information Systems Laboratory. This work was supported by NSF Grant 1624414-0.
REFERENCES
Albers, J. R., and T. R. Nathan, 2012: Pathways for communicating the effects of stratospheric ozone to the polar vortex: Role of zonally asymmetric ozone. J. Atmos. Sci., 69, 785–801, https://doi.org/10.1175/JAS-D-11-0126.1.
Afiesimama, E. A., 2007: Annual cycle of the mid-tropospheric easterly jet over West Africa. Theor. Appl. Climatol., 90, 103–111, https://doi.org/10.1007/s00704-006-0284-y.
Bercos-Hickey, E., T. R. Nathan, and S.-H. Chen, 2017: Saharan dust and the African easterly jet–African easterly wave system: Structure, location, and energetics. Quart. J. Roy. Meteor. Soc., 143, 2797–2808, https://doi.org/10.1002/qj.3128.
Burpee, R. W., 1972: The origin and structure of easterly waves in the lower troposphere of North Africa. J. Atmos. Sci., 29, 77–90, https://doi.org/10.1175/1520-0469(1972)029<0077:TOASOE>2.0.CO;2.
Burpee, R. W., 1974: Characteristics of North African easterly wave during the summers of 1968 and 1969. J. Atmos. Sci., 31, 1556–1570, https://doi.org/10.1175/1520-0469(1974)031<1556:CONAEW>2.0.CO;2.
Burpee, R. W., 1975: Some features of synoptic-scale wave based on a compositing analysis of GATE data. Mon. Wea. Rev., 103, 921–925, https://doi.org/10.1175/1520-0493(1975)103<0921:SFOSWB>2.0.CO;2.
Chen, S.-H., S.-H. Wang, and M. Waylonis, 2010: Modification of Saharan air layer and environmental shear over the eastern Atlantic Ocean by dust-radiation effects. J. Geophys. Res., 115, D21202, https://doi.org/10.1029/2010JD014158.
Chen, S.-H., Y.-C. Liu, T. R. Nathan, C. Davis, R. Torn, N. Sowa, C.-T. Cheng, and J.-P. Chen, 2015: Modeling the effects of dust-radiative forcing on the movement of Hurricane Helene (2006). Quart. J. Roy. Meteor. Soc., 141, 2563–2570, https://doi.org/10.1002/qj.2542.
Chou, M. D., and M. J. Suarez, 1999: A solar radiation parameterization for atmospheric studies. NASA Tech. Memo. 104606, Vol. 15, 40 pp., https://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/19990060930.pdf.
Chou, M. D., M. J. Suarez, X. Z. Liang, and M. M. H. Yan, 2001: A thermal infrared radiation parameterization for atmospheric studies. NASA Tech. Memo. 104606, Vol. 19, 102 pp., https://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/20010072848.pdf.
Dezfuli, A. K., and S. E. Nicholson, 2011: A note on long-term variations of the African easterly jet. Int. J. Climatol., 31, 2049–2054, https://doi.org/10.1002/joc.2209.
Grist, J. P., and S. E. Nicholson, 2001: A study of the dynamic factors influencing the rainfall variability in the West African Sahel. J. Climate, 14, 1337–1359, https://doi.org/10.1175/1520-0442(2001)014<1337:ASOTDF>2.0.CO;2.
Grist, J. P., S. E. Nicholson, and A. I. Barcilon, 2002: Easterly waves over Africa. Part II: Observed and modeled contrasts between wet and dry years. Mon. Wea. Rev., 130, 212–225, https://doi.org/10.1175/1520-0493(2002)130<0212:EWOAPI>2.0.CO;2.
Grogan, D. F. P., and C. D. Thorncroft, 2019: The characteristics of African easterly waves coupled to Saharan mineral dust aerosols. Quart. J. Roy. Meteor. Soc., 145, 1130–1146, https://doi.org/10.1002/qj.3483.
Grogan, D. F. P., T. R. Nathan, and S.-H. Chen, 2016: Effects of Saharan dust on the linear dynamics of African easterly waves. J. Atmos. Sci., 73, 891–911, https://doi.org/10.1175/JAS-D-15-0143.1.
Grogan, D. F. P., T. R. Nathan, and S.-H. Chen, 2017: Saharan dust and the nonlinear evolution of the African easterly jet–African easterly wave system. J. Atmos. Sci., 74, 27–47, https://doi.org/10.1175/JAS-D-16-0118.1.
Hess, M., P. Koepke, and I. Shult, 1998: Optical properties of aerosols and clouds: The software package OPAC. Bull. Amer. Meteor. Soc., 79, 831–844, https://doi.org/10.1175/1520-0477(1998)079<0831:OPOAAC>2.0.CO;2.
Hosseinpour, F., and E. M. Wilcox, 2014: Aerosol interactions with African/Atlantic climate dynamics. Environ. Res. Lett., 9, 075004, https://doi.org/10.1088/1748-9326/9/7/075004.
Hsieh, J.-S., and K. H. Cook, 2007: A study of the energetics of African easterly waves using a regional climate model. J. Atmos. Sci., 64, 421–440, https://doi.org/10.1175/JAS3851.1.
Jones, C., N. Mahowald, and C. Luo, 2003: The role of easterly waves on African desert dust transport. J. Climate, 16, 3617–3628, https://doi.org/10.1175/1520-0442(2003)016<3617:TROEWO>2.0.CO;2.
Jones, C., N. Mahowald, and C. Luo, 2004: Observational evidence of African desert dust intensification of easterly waves. Geophys. Res. Lett., 31, L17208, https://doi.org/10.1029/2004GL020107.
Jury, M. R., and M. J. Santiago, 2010: Composite analysis of dust impacts on African easterly waves in the Moderate Resolution Imaging Spectrometer era. J. Geophys. Res., 115, D16213, https://doi.org/10.1029/2009JD013612.
Knippertz, P., and M. C. Todd, 2010: The central west Saharan dust hot spot and its relation to African easterly waves and extratropical disturbances. Geophys. Res. Lett., 115, D12117, https://doi.org/10.1029/2009JD012819.
Kok, J. F., 2011: A scaling theory for the size distribution of emitted dust aerosols suggests climate models underestimate the size of the global dust cycle. Proc. Natl. Acad. Sci. USA, 108, 1016–1021, https://doi.org/10.1073/pnas.1014798108.
Kwon, H. J., 1989: A reexamination of genesis of African waves. J. Atmos. Sci., 46, 3621–3631, https://doi.org/10.1175/1520-0469(1989)046<3621:AROTGO>2.0.CO;2.
Lavaysse, C., J.-P. Chaboureau, and C. Flamant, 2011: Dust impact on West African heat low in summertime. Quart. J. Roy. Meteor. Soc., 137, 1227–1240, https://doi.org/10.1002/qj.844.
Ma, P.-L., K. Zhang, J. J. Shi, T. Matsui, and A. Arking, 2012: Direct radiative effect of mineral dust on the development of African easterly waves in late summer, 2003–07. J. Appl. Meteor. Climatol., 51, 2090–2104, https://doi.org/10.1175/JAMC-D-11-0215.1.
Nathan, T. R., D. F. P. Grogan, and S.-H. Chen, 2017: Subcritical destabilization of African easterly waves by Saharan mineral dust. J. Atmos. Sci., 74, 1039–1055, https://doi.org/10.1175/JAS-D-16-0247.1.
Nathan, T. R., D. F. P. Grogan, and S.-H. Chen, 2019: Saharan dust transport during the incipient growth phase of African easterly waves. Geosciences, 2019, 9, 388, https://doi.org/10.3390/geosciences9090388.
Newell, R. E., and J. W. Kidson, 1984: African mean wind changes between Sahelian wet and dry periods. Int. J. Climatol., 4, 27–33, https://doi.org/10.1002/joc.3370040103.
Norquist, D. C., E. E. Recker, and R. J. Reed, 1977: The energetics of African wave disturbances as observed during phase III of GATE. Mon. Wea. Rev., 105, 334–342, https://doi.org/10.1175/1520-0493(1977)105<0334:TEOAWD>2.0.CO;2.
Paradis, D., J.-P. Lafore, J.-L. Redelsperger, and V. Balaji, 1995: African easterly waves and convection. Part I: Linear simulations. J. Atmos. Sci., 52, 1657–1679, https://doi.org/10.1175/1520-0469(1995)052<1657:AEWACP>2.0.CO;2.
Pytharoulis, I., and C. D. Thorncroft, 1999: The low-level structure of African easterly waves in 1995. Mon. Wea. Rev., 127, 2266–2280, https://doi.org/10.1175/1520-0493(1999)127<2266:TLLSOA>2.0.CO;2.
Quijano, A., I. N. Sokolik, and O. B. Toon, 2000: Radiative heating rates and direct radiative forcing by mineral dust in cloudy atmospheric conditions. J. Geophys. Res., 105, 12 207–12 219, https://doi.org/10.1029/2000JD900047.
Reale, O., K. M. Lau, and A. da Silva, 2011: Impact of interactive aerosol on the African easterly jet in the NASA GEOS-5 Global Forecasting System. Wea. Forecasting, 26, 504–519, https://doi.org/10.1175/WAF-D-10-05025.1.
Reed, R. J., D. C. Norquist, and E. E. Recker, 1977: The structure and properties of African wave disturbances as observed during phase III of GATE. Mon. Wea. Rev., 105, 317–333, https://doi.org/10.1175/1520-0493(1977)105<0317:TSAPOA>2.0.CO;2.
Reed, R. J., E. Klinker, and A. Hollingsworth, 1988: The structure and characteristics of African easterly wave disturbances as determined from the ECMWF operational analysis/forecast system. Meteor. Atmos. Phys., 38, 22–33, https://doi.org/10.1007/BF01029944.
Rennick, M. A., 1976: The generation of African waves. J. Atmos. Sci., 33, 1955–1969, https://doi.org/10.1175/1520-0469(1976)033<1955:TGOAW>2.0.CO;2.
Rennick, M. A., 1981: Some sensitivity experiments with an African wave model. J. Atmos. Sci., 38, 106–113, https://doi.org/10.1175/1520-0469(1981)038<0106:SSEWAA>2.0.CO;2.
Simmons, A. J., and B. J. Hoskins, 1976: Baroclinic instability on the sphere: Normal modes of the primitive and quasi-geostrophic equations. J. Atmos. Sci., 33, 1454–1477, https://doi.org/10.1175/1520-0469(1976)033<1454:BIOTSN>2.0.CO;2.
Thorncroft, C. D., and B. J. Hoskins, 1994: An idealized study of African easterly waves. I: Linear theory. Quart. J. Roy. Meteor. Soc., 120, 953–982, https://doi.org/10.1002/qj.49712051809.
Thorncroft, C. D., 1995: An idealized study of African easterly waves. III: More realistic basic states. Quart. J. Roy. Meteor. Soc., 121, 1589–1614, https://doi.org/10.1002/qj.49712152706.
Tompkins, A. M., C. Cardinali, J.-J. Morcette, and M. Rodwell, 2005: Influence of aerosol climatology on forecasts of the African easterly jet. Geophys. Res. Lett., 32, L10801, https://doi.org/10.1029/2004GL022189.
Yu, H., M. Chin, D. M. Winker, A. H. Omar, Z. Liu, C. Kittaka, and T. Diehl, 2010: Global view of aerosol vertical distributions from CALIPSO lidar measurements and GOCART simulations: Regional and seasonal variations. J. Geophys. Res., 115, D00H30, https://doi.org/10.1029/2009JD013364.
The two pathways are not unique to dust; they operate in any system that involves the interaction between the circulation and a trace shortwave radiative absorber, such as stratospheric ozone (cf. Albers and Nathan 2012).
For a wave propagating in a zonal-mean current, the Doppler-shifted frequency is defined as