1. Introduction and background
The West African Sahel lies within the semiarid to arid transition zone between the wet and humid equatorial zone of Africa to the south and the Sahara Desert to the north (Fig. 1). This topography creates a significant north–south precipitation gradient and a climate with highly variable precipitation on a number of spatial and temporal scales. Evapotranspiration from the land surface and the horizontal transport of humid air from the monsoon are key processes that can amplify or attenuate the precipitation distribution and intensity throughout the wet season in the Sahel.
The dynamics of the West African monsoon are governed mainly by the meridional distribution of boundary layer entropy (Eltahir and Gong 1996). Meridional potential temperature profiles in the boundary layer delineate the location and intensity of the African easterly jet (Thorncroft and Blackburn 1999). This dynamical setup provides West African precipitation with an increased sensitivity to conditions at the surface (Cook 1999). Hodges and Thorncroft (1997) have shown that African convective complexes exhibit a general westward movement, and an average lifetime of approximately one day within the Sahel. This allows convective systems traveling through a region to then be strengthened or weakened by interactions with the land surface. This study investigates the effect of soil moisture on the potential for deep convection in a semiarid environment in West Africa.
In arid and semiarid regions, vegetation cover and soil moisture are key components in the partitioning of energy at the land surface into sensible and latent heat fluxes. The Sahel consists of diverse land covers ranging from dense woody savannas and grasslands in the south to sparsely vegetated semidesert in its northernmost regions. Mohr et al. (2003) note appreciable differences in the water budgets of simulated convective lines and the spatial and temporal patterns of resulting precipitation over three different land covers within West Africa. The amount of vegetation covering the land surface has been shown to influence precipitation amounts in the Sahel, and a reduction in coverage can promote drought conditions (Xue and Shukla 1993; Taylor et al. 2002).
In some portions of the Sahel, vegetation often grows in dense strips intertwined with large areas of completely bare soil. This vegetation pattern is referred to as “tiger bush,” and these areas are thought to have an important influence on the hydrology of the Sahel (Wallace and Holwill 1997). The tiger bush land cover is common throughout the Sahel and occupies about one-third of Sahelian Niger (Ambouta 1984). The alternating strips of bare ground and mixed vegetation are aligned approximately perpendicular to the height gradient of the gently sloping terrain. The 10−30-m-wide strips of vegetation occupy approximately 20%–35% of the land cover (Wallace and Holwill 1997). Bromley et al. (1997) indicate that the large bare-soil areas in this environment act to recharge the groundwater supply under the bands of vegetation by allowing a significant amount of runoff generated from the large bare-soil strips to flow into the more permeable surface under the vegetation cover. The role of the bare-soil area as rainfall collectors for the vegetated strips and their effects on groundwater recharge has been the focus of numerous studies within of the tiger bush environment. Detailed studies of this nature can be found in Galle et al. (1999), Valentin and d’Herbes (1999), and Wu et al. (2000).
The spatial distribution of soil moisture also plays an important role in determining precipitation amounts in the Sahel. Betts and Ball (1995, 1998) show that given similar soil temperatures and net radiation, wet soil moisture conditions generate larger latent heat fluxes, increasing the amount of low-level moist entropy and favoring the development of deep convection. Taylor et al. (1997a,b) and Taylor and Lebel (1998) investigated the variability in the boundary layer and its relationship to mesoscale gradients in soil moisture during the Hydrologic Atmospheric Pilot Experiment in the Sahel (HAPEX-Sahel) [experiment details in Goutorbe et al. (1994)]. They found that mesoscale soil moisture gradients induced boundary layer anomalies of moist static energy (MSE), and this produced a positive feedback with subsequent convective lines in several locations throughout the field experiment. Findell and Eltahir (2003) examine the problem of how the early morning atmospheric thermodynamic structure affects the interaction between wet and dry soil moisture environments and ultimately convective initiation. They develop a framework to assist in determining if precipitation within a particular region exhibits a long-term feedback with soil moisture and describe atmospheric structures where wet soils are more likely to trigger precipitation than dry soils.
Coupled land–atmosphere models are a useful hydrometeorological tool for evaluating the impact the underlying land cover and soil moisture have on spatial and temporal variability of precipitation. Two individual cases will be modeled in the late summer of 1992 using data from the HAPEX-Sahel field experiment. The goal of this study is not to simulate a particular event but to identify differences in boundary layer growth, convective initiation and intensity, and precipitation within wet and dry soil moisture regimes in the tiger bush environment.
2. Data and methodology
a. Numerical model
A two-dimensional (2D) coupled land–atmosphere (cloud resolving) numerical model with open lateral boundary conditions is used for all of the simulations performed in this work. Prior studies have shown that the time-mean kinematic and thermodynamic properties of convective lines are well simulated in 2D models, and they have the advantage of computational speed and efficiency when many simulations are required, such as in a sensitivity study (Nicholls et al. 1988; Ferrier et al. 1996; Tao et al. 1996; Xu and Randall 1996; Grabowski et al. 1998; Lucas et al. 2000; Mohr et al. 2003). The Goddard Cumulus Ensemble (GCE) is a nonhydrostatic, anelastic, numerical cloud-resolving model that includes variables for horizontal and vertical velocities, potential temperature, water vapor mixing ratio, and turbulent kinetic energy (Tao and Simpson 1993). The subgrid-scale turbulence used in the GCE model is based on work by Klemp and Wilhelmson (1978). In their approach, one prognostic equation is solved for subgrid turbulent kinetic energy (TKE), which is then used to specify the eddy coefficients. The effect of condensation on the generation of subgrid-scale kinetic energy is also incorporated in the model (Soong and Ogura 1980). The cloud microphysics include a parameterized two-category liquid water scheme (cloud water and rain), and a parameterized three-category ice-phase scheme (cloud ice, snow, and hail/graupel; Lin et al. 1983; Rutledge and Hobbs 1984). Solar and infrared radiation parameterizations are also included in the model (Tao et al. 1996).
In this study, GCE is coupled to the Parameterization for Land–Atmosphere–Cloud Exchange (PLACE) (Wetzel and Boone 1995). The PLACE model is a detailed, interactive process model of the heterogeneous land surface (soil and vegetation) and adjacent near-surface atmosphere. PLACE emphasizes the vertical transport of moisture and energy through a five-layer soil moisture and a seven-layer soil temperature column to the overlying surface. The additional two soil temperature layers are used to aid in resolving large temperature gradients near the surface. Momentum, sensible, and latent heat fluxes are calculated using similarity relationships (Zilitinkevich 1975; Businger et al. 1971). In PLACE, each model grid cell contains a specified fraction of fully vegetated surface, with the rest of the surface treated as nonvegetated bare soil. Separate evapotranspiration equations are used to compute supply- and demand-limited rates for vegetation and bare-soil areas. The demand-limited rate is calculated using the mixing-ratio gradient between the ground and the surface layer and the magnitude of the canopy resistance (Noilhan and Planton 1989). Supply-limited rates are dependent upon the gradient between the water potentials of the soil and plant (Wetzel and Boone 1995). The resulting latent heat flux in a grid cell is then computed as a weighted sum of the fluxes from the vegetated and bare-soil areas. The exchanges of radiation and heat, momentum, and moisture fluxes couple the GCE and PLACE models (Tao et al. 2003). This same two-dimensional version of GCE–PLACE has previously been used to investigate both landscape-generated deep convection (Lynn et al. 1998) and the sensitivity of West African convective line water budgets to land cover (Mohr et al. 2003).
A horizontal domain of approximately 1000 km divided into 2048 grid points with 1990 inner points at 0.5-km spacing and a stretched (1:1.06) grid on either side is used in all simulations. This domain consists of only tiger bush vegetation and does not represent the entire area outlined in Fig. 1, as this area contains several other land-cover types. The model consists of 33 grid points in the vertical starting at 0.0 km and ending at 21.5 km. This grid was also stretched with highest resolution in the lowest levels with the spacing between levels ranged from 80 m at the surface to 1200 m at the top of the troposphere. GCE operated on a 5-s time step and called PLACE for surface fluxes at 3-min intervals. All simulations started at local midnight and ran for 24 h, allowing for the examination of a full diurnal cycle. No lifting mechanism, such as a warm bubble or cold pool, was applied so that the initial atmospheric structure could evolve over the course of the day.
b. Initial conditions and numerical experiments
The vegetation, soil type, soil moisture, and soil temperature data used to initialize the model were derived from the hydrologic HAPEX-Sahel field experiment. HAPEX-Sahel was conducted during 1991–92 within a 1° by 1° square in southwestern Niger (2°–3°E and 13°–14°N). The goal of HAPEX-Sahel was to improve the parameterizations of land surface–atmosphere interactions at the global circulation model scale, because of the hypothesized link between land degradation and climatic changes in the Sahel. The experiment combined remote sensing and ground-based measurements and contained an intensive observation period that captured the transition of the wet season to the dry season in August through October of 1992. All of the land surface data originated from within the southern supersite within the HAPEX-Sahel study area during the intensive observation period. This site was predominantly composed of the tiger bush land cover but also contained areas of millet, grassy savanna, and fallow bush. A full description of the site and the experimental methods therein can be found in Wallace et al. (1994).
The tiger bush land cover is deeply rooted, highly reflective, sparsely vegetated, and able to withstand the highly oxidizing and leaching Sahel climate. The land cover in the study simulations was specified so that on average ∼67% was bare crusted soil and the rest vegetation (based on aerial photographs from the experiment site). The vegetated strips within the tiger bush land cover are about 10–30 m wide by 100–300 m long and consist primarily of 2–4–m-tall shrubs (Guiera Senegalensis and Combretum micranthum) (Wallace and Holwill 1997). There exists very little herbaceous growth with the exception of the upslope edge of the shrubs. A large portion of the bare-soil transects in the tiger bush land cover are strongly crusted and have very low infiltration rates and very low saturated hydraulic conductivity (Bromley et al. 1997). Below the surface crusting, the soil texture is principally sandy loam (Wallace and Holwill 1997). Hence, the soil type used in this study is a sandy loam soil.
Initial soil moisture conditions define the simulations used in this study. A wet regime refers to a simulation that contains a wet soil moisture profile and a dry regime refers to a simulation that contains a dry soil moisture profile. Table 1 lists the initial soil moisture profiles and parameters, and Table 2 lists the vegetation parameters used in both numerical case studies. Soil moisture, soil temperature, and vegetation cover were randomly varied across the model domain to account for the natural variability of the tiger bush land cover. In addition, the saturated hydraulic conductivity of the first soil moisture layer was decreased in order to account for the strongly crusted uppermost soil layer within PLACE. The wet soil moisture profile is typical of a soil column that has been exposed to several days of consecutive precipitation, whereas the dry soil moisture profile is more representative of soils that have been drying for approximately one week. This was based on observations of soil transects within the southern supersite. Initial volumetric soil moisture in the wet regime ranges from 0.26 cm3 cm−3 (65% saturation) in the lowest layers to as little as 0.14 cm3 cm−3 (35% saturation) in the near-surface layers. Initial volumetric soil moisture in the dry regime ranges from 0.15 cm3 cm−3 (38% saturation) in the lowest layers to as little as 0.04 cm3 cm−3 (10% saturation) in the near-surface layers.
The sounding data used to initialize the numerical simulations came from field observations during HAPEX-Sahel and operational rawinsonde launches from Niamey, Niger. The field observations were used in conjunction with the operational soundings to make complete and detailed initial soundings for the simulations. The initial soundings were then smoothed to remove any sharp changes in wind speed, wind direction, and humidity to prevent numerical instability within the GCE model.
Before attempting a 24-h simulation, the atmosphere and the land surface must be in quasi equilibrium in order to avoid generating large nonphysical surface fluxes. First-guess soil moisture and temperature profiles were created, and PLACE was run offline forced by actual surface observations in the southern supersite of pressure, temperature, wind speed, specific humidity, and radiation from midnight to 0300 LST. The resulting soil moisture and temperature profiles at the end of the first 3-h offline simulation became the initial soil moisture and temperature profiles for a repeat of the 3-h simulation. The 3-h simulation was repeated (about 20 iterations) until the initial and final soil moisture profiles were nearly identical (soil moisture within ±0.001 cm3 cm−3 and soil temperature within ±0.1 K). The resulting initial soil moisture values are listed in Table 1. This methodology was applied to both soil moisture profiles within each numerical case study so that each regime (wet and dry) had its own initial soil temperature profile. By changing only the initial soil moisture profile, we can isolate the physical processes important in boundary layer growth and convective initiation within the tiger bush environment.
Two numerical case studies are presented in this paper, a convective case with precipitation, 25 August 1992, and a fair weather case, 1 September 1992. The objectives of the fair weather case study are to analyze how much wet soil might increase the potential for convective initiation and how much dry soil might reinforce fair weather conditions. The fair weather and convective cases will be compared to diagnose the influences of the different soil moisture regimes within these environments.
3. 25 August 1992 case study results
a. Initial atmospheric conditions
The 0000 UTC 25 August 1992 sounding is depicted in Fig. 2. This sounding has several key attributes. The environment is unstable with an initial convective available potential energy (CAPE) value of 1860 J kg−1 and convective inhibition (CIN) of 80 J kg−1. Both CAPE and CIN were calculated by lifting the average of the lowest 50 hPa.
The core of the African easterly jet is located around 630 hPa, and the core of the tropical easterly jet is located near 180 hPa (Fig. 2b). The location and strength (15 m s−1) of the African easterly jet in this environment creates significant low-level wind shear. There is approximately −5.7 × 10−3 s−1 of shear from height of the maximum low-level wind at 0.47 km to 3.65 km (location of the African easterly jet). The meridional component of the wind is southerly from the surface to about 900 hPa, indicating strong moist monsoonal flow at this time.
b. Surface fluxes
In both the wet and dry regimes, the sensible heat flux was larger than the latent heat flux. The sensible heat flux in the dry regime was nearly 50% higher than the sensible heat flux in the wet regime. Conversely, the wet regime had 60% more latent heat flux than the dry regime. In both regimes, the latent heat flux peaks in the morning were associated with maxima in evapotranspiration. In the afternoon, evaporating rainfall at the surface was the principal cause, allowing latent heat flux to continue after sunset.
c. Boundary layer development
Equivalent potential temperature (θe) is a proxy for identifying the distribution of MSE within each regime. Figure 4 is the distribution of θe of the lowest 3 km in the model at hour 13, just prior to the triggering of deep convection. The wet regime has a 1-km layer of air with θe greater than 342 K, and almost the entire surface layer exceeds 344 K. In both of these layers, the dry regime θe is at least 2 K cooler than the wet regime. In the surface layer the dry regime θe does not exceed 344 K. The boundary layer depth is defined as the depth of the well-mixed layer. In Fig. 5, the boundary layer in the wet regime is shallower (1.39 km versus 1.78 km) and exhibits a stronger gradient of θe (MSE) between 1 and 3 km in the early afternoon. The evapotranspiration in the wet regime produced a much larger latent heating term (Lw), despite a smaller sensible heating term (CpT), than the dry regime. The boundary layer of the dry regime grew faster than the boundary layer of the wet regime and thus was more prone to the entrainment of drier air from the free atmosphere (Fig. 4). The increased entrainment of drier air in the dry regime boundary layer contributed to its weaker vertical gradient of MSE throughout the simulation.
Differences in boundary layer growth and development led to differences between the two regimes in the temperature and moisture stratification of the free atmosphere. In Fig. 6, the CAPE drops early in both regimes due to the formation of a small radiation inversion at the surface. The CAPE then builds up in the morning until the onset of convection, which then erodes the CAPE values throughout the afternoon. By noon, the wet regime had 2220 J kg−1 of CAPE (Fig. 6a), while the dry regime (Fig. 6b) had 1530 J kg−1 (32% less). The slower-growing boundary layer of the wet regime was less prone to the entrainment of drier air aloft. In the warmer, moister, shallower, wet regime boundary layer (Fig. 5a), significant CAPE was produced and remained higher than in the dry regime (Fig. 5b) from midmorning onward. In Figs. 5 and 6, the wet regime appears more capable of sustaining deep convection with heavy precipitation than the dry regime.
d. Convection, precipitation, and final water budgets
Enhanced cumulus began forming around hour 13 in the dry regime and hour 14 in the wet regime. Figure 7 is the surface precipitation rate recorded at every 5 min across the entire model domain. Convection and precipitation in the dry regime started slightly earlier (hour 13.5 versus hour 14). Faster boundary layer growth in the dry regime lifted parcels to their level of free convection sooner. However, all precipitation ended around hour 18. Most of the convective lines in both regimes lasted no longer than 1 h. By using the term “convective lines,” we are referring to lines of cumulonimbus clouds in space and time as depicted in Fig. 7 and not lines in space only. Fewer convective lines developed in the wet regime, but several lasted over 2 h. Both regimes exhibited a second onset of convection between hours 16 and 17. Once initiated, the wet regime supported convection until the end of the simulation whereas all but one of the convective lines in the dry regime lasted less than 2 h.
The duration, spatial distribution, and intensity of convection varied between the two regimes. Figure 8 is a time series of domain maximum updrafts for both regimes. The peak vertical velocity in the wet regime was 1 m s−1 lower than the dry regime, but it maintained strong updrafts (above 20 m s−1) until the end of the simulation. Updrafts above 10 m s−1 ended between hour 16 and 17 in the dry regime. The differences in updrafts revealed that convection was generally more intense in the wet regime, producing differences in the final precipitation rate distribution (Fig. 9). The wet regime has a higher number of counts in all of the categories, especially the >50 mm h−1 bin, and had 77 counts over 100 mm h−1 compared to 9 counts in the dry regime. Table 3 is the frequency distribution of surface precipitation rates from Fig. 9. The wet regime exhibits a higher percentage of precipitation rates in all of the binned categories except for the 0–5-mm bin where the dry regime has slightly over 3% more in this category. Noteworthy in this table is the higher percentage of precipitation rates falling within the largest precipitation rates bin. The percentage of precipitation rates within the wet regime is more than double those produced in the dry regime, highlighting the more intense precipitation produced in the wet regime.
The distribution of precipitation rates in the wet regime is the result of more intense, longer-lasting convection, arising from the manner in which the soil moisture influenced boundary layer growth and development. High sensible heat flux and significant entrainment into the dry regime boundary layer resulted in convection that was typically weak and short lived.
4. 1 September 1992 case study results
a. Initial atmospheric conditions
The objective of this case study is to observe how the different soil moisture regimes alter fair weather conditions and which regime would favor more intense convection should preexisting disturbances such as a gust front from a transient squall line perturb the environment. In addition, we reference results from the previous case study to assess the impact of the land surface in convective versus fair weather environments. The 0000 UTC sounding used to initialize the wet and dry regime simulations is depicted in Fig. 10. This sounding is typical of a fair weather sounding within the tiger bush environment. The lower-tropospheric subsidence inversion and associated dry air intrusion at 850 hPa generate a significant amount of CIN, 250 J kg−1, in this environment. The CAPE of this sounding is only 470 J kg−1, and coupled with the large amount of CIN, creates a stable environment.
The core of the African easterly jet is located around 650 hPa and the core of the tropical easterly jet is located near 170 hPa. There are −6.2 × 10−3 s−1 of shear from the height of the maximum low-level wind at 0.5 to 3.5 km (location of the African easterly jet). This value of shear is higher than calculated in the convective case study and is due to the lower altitude of the African easterly jet and stronger southwesterly flow near the surface. The meridional component of the wind is southerly to 880 hPa yielding a slightly deeper monsoon flow deeper than the convective case.
b. Surface fluxes
The time series of domain-averaged surface fluxes for both regimes are in Fig. 11. The wet regime had slightly more net radiation than the dry regime. As in the previous case study, this was from a combination of cooler soil temperatures (up to 5 K) and the lower albedo of the wetter soils. The wet regime exhibited a 50% larger ground heat flux, due to the higher thermal conductivity of the wetter soils resulting in 13% less available energy than in the dry regime. Increased evapotranspiration in the wet regime produced nearly 3 times the latent heat flux of the dry regime. However, unlike the previous case study, the latent heat flux peaks only in the morning hours because there is no evaporating rainfall in the late afternoon. The sensible heat flux in the dry regime peaked about 200 W m−2 higher than in the wet regime. This case study had similar amounts of net radiation, ground heat flux, and available energy compared to the wet and dry regimes of the previous case study. The latent heat flux in the wet regime was nearly 22% larger than the previous case study and was produced by a larger surface vapor pressure deficit in the fair weather environment, making higher evapotranspiration rates from the tiger bush land cover possible. In the dry regime latent heat flux was 20% less than in the convective case study because of warmer near-surface air temperatures. Higher air temperatures increased the stomatal resistance of the vegetation in PLACE thereby reducing transpiration. The sensible heat flux in the wet regime was similar to the convective case study, but 11% higher in the dry regime due to a warmer (1–2 K) surface temperature.
c. Boundary layer development
The differences in surface fluxes between the two regimes created two different boundary layers. In Fig. 12a, the wet regime develops a stronger vertical gradient in θe (MSE) and, since no convection develops, maintains it through the end of the simulation. The vertical gradient of MSE in the wet regime in Fig. 12a is stronger than the wet regime in the convective case (Fig. 6a). The boundary layer in the dry regime (Fig. 12b) entrained copious dry air, producing a weaker vertical gradient of MSE. The dry regime in this case study behaved similarly to the dry regime in the convective case study, as both were prone to the entrainment of drier air aloft. There is a more obvious difference in the vertical gradients of MSE between the wet and dry regimes in this case study compared to the convective case study.
The CAPE in the dry regime increased to 700 J kg−1 by noon, falling to 20 J kg−1 by the end of the simulation (Fig. 13). The wet regime, where entrainment was not as prevalent, increased to 1230 J kg−1 of CAPE by noon, falling to 780 J kg−1 by the end of the simulation. This is consistent with the previous case study in which the wet regime builds up significantly more CAPE than the dry regime. The fair weather wet regime builds up less CAPE, but since there is no convection to eliminate CAPE, there is more remaining at the end of the simulation.
5. Discussion
a. Wet versus dry soil moisture regimes
Surface fluxes within GCE-PLACE have been presented for all of the simulations performed in this study. Compared to observations during HAPEX-Sahel, GCE-PLACE underestimates the overall evaporative fraction [E/(H + E)] of the tiger bush land cover. However, the relative increases in evaporative fraction between wet soil moisture days and dry soil moisture days were consistent with observations. The differences in evapotranspiration from the tiger bush land cover between the two regimes proved to be responsible for differences in boundary layer growth and development in both fair weather and convective cases. The wet regimes produced shallower boundary layers with stronger vertical gradients of MSE (θe), more MSE per unit area, and more CAPE throughout the simulation, particularly in the 1 September fair weather case study. The increased sensible heat flux in the dry regime boundary layers resulted in faster boundary layer growth and, in the convective case, more entrainment of dry air into rising parcels. In the 25 August convective case, dry regime parcels reached their levels of free convection faster than in the wet regime, initiating deep convection and precipitation earlier in the afternoon.
More convective lines developed in the dry regime but they were generally less intense and short lived compared to the convective lines in the wet regime. For the dry regime, the combined effect of numerous convective cells and increased entrainment into the boundary layer air quickly eroded the vertical gradient of MSE (θe) and reduced the CAPE, making it impossible to sustain convection throughout the evening hours. In the wet regime, the boundary layer was able to sustain deep convection until the end of the simulation, producing 2.25 times more precipitation.
b. Fair weather cases versus convective cases
The vertical gradient in vapor pressures at the surface and in the surface layer is a key component of the calculation of evapotranspiration in PLACE. The fair weather case sounding (Fig. 10a) was drier than the convective case sounding (Fig. 2a) and capped by a strong subsidence inversion at 850 hPa. Such a situation would have the potential to produce high evapotranspiration rates should enough soil moisture be available. There was slightly more latent heat flux in the wet regime in the fair weather case compared to the wet regime in the convective case, but similar sensible heat fluxes. The biggest difference in the surface fluxes is between the dry regimes. Maximum sensible heat flux is 20% higher, because of warmer soils and in turn warmer skin temperatures, in the dry regime of the fair weather case, producing a higher boundary layer (Figs. 5b and 12b). Despite the height of the fair weather dry regime boundary layer, it was still well below the level of free convection determined by the fair weather lapse rate.
Differences in surface fluxes between the two regimes were more pronounced in the fair weather case. Because of a stronger surface vapor pressure gradient, the difference in latent heat flux between the wet and dry regimes was 11% higher in the fair weather case than in the convective case. The dry regime in the fair weather case produced faster boundary layer growth from the morning into the early afternoon compared to the dry regime in the convective case. While there was more entrainment in the dry regimes than in the wet regimes in both cases, the boundary layer in the dry regime in the fair weather case entrained more warm, dry air than did the dry regime in the convective case. The dry regime in the fair weather case exhibited a weaker vertical gradient of MSE than the dry regime in the convective case. This resulted in bigger differences between the wet and dry soil moisture regimes in the fair weather case with respect to MSE gradients. In the wet regimes of both cases, the surface fluxes and thus the boundary layer growth were comparable. The CIN and capping inversion in the fair weather sounding made it impossible for rising parcels to reach their level of free convection even in the wet regime. Dry entrainment in the fair weather simulations was responsible for making even shallow clouds rare.
The different soil moisture regimes had a noticeable impact in both case studies, although the initial sounding was the most important variable determining a fair day versus a rainy day. Without a moist southwesterly monsoon flow, precipitation is unlikely in the tiger bush environment even when the soil is wet. That is, land surface processes can modify the development of convection but cannot solely determine convective initiation. The wet regime of the fair weather case presents an interesting situation. The high CAPE of the late afternoon (Fig. 13a) could make it possible for a preexisting disturbance, such as a westward-moving squall line, to trigger new convection and precipitation. It is easier to imagine a squall line meeting its demise in the dry regime environment (Fig. 13b) than creating new cells. For areas such as central Niger where fair weather soundings are common, the results of the fair weather case study suggest land–atmosphere cooperation mechanisms that could produce the mesoscale soil moisture gradients that Taylor and Lebel (1998) observed during HAPEX-Sahel.
Our simulated convective lines are distinctly local, representative of small convective systems that develop from daytime heating of an unstable boundary layer during an active monsoon (Mohr et al. 2003). From TRMM data, Mohr (2004) and Mohr and Thorncroft (2006) observe that a majority of West African convective systems are small, weak, develop near locations of significant orography during active monsoons, and have a life cycle of a couple of hours. The results of the convective case in this study appear to describe the land–atmosphere exchange processes responsible for the development of the convective systems described in the observational studies. The differences between the fair weather case and convective case highlight the importance of the land surface in fair weather conditions via significant differences in CAPE generation between the two soil moisture regimes. Should an organized large convective system interact with a fair weather environment similar to the one presented here, it is reasonable to hypothesize that the system would exhibit comparable sensitivity to the soil moisture conditions.
6. Conclusions
Using data from the HAPEX-Sahel field experiment, a tiger bush land cover was specified within the GCE–PLACE model along with two different soil moisture profiles. A wet regime classified a simulation that contained a wet soil moisture profile. This profile was representative of soils that received several days of precipitation. A dry regime classified a simulation that contained a dry soil moisture profile. This profile was representative of soils that have been drying out for approximately one week. Two numerical case studies using different atmospheric soundings were presented, one case a convective precipitation event and the other case having fair weather. The different soil moisture regimes proved to be important to the development of the planetary boundary layer in both cases.
The wet regimes in all of the case studies had more net radiation, and a higher latent heat flux, than the dry regimes. The increased latent heat flux in the wet regimes resulted in moister and shallower boundary layers with less entrainment of drier air aloft. This allowed the wet regimes to create and maintain stronger vertical gradients of MSE and higher CAPE values throughout the simulation. In both case studies, the wet regime created a boundary layer that was more favorable to deep convection than the dry regime. In the convective case, precipitation began slightly earlier in the dry regime, but its convective lines were less intense and shorter lived than in the wet regime due to increased entrainment into the dry regime boundary layer. Convective lines in the dry regime generated approximately 55% less precipitation than the wet regimes.
The differences in surface fluxes between regimes in the fair weather case were slightly more pronounced than in the convective case study. These differences had noticeable impacts on boundary layer growth and development between the individual fair weather cases. The contrast in the evolution of the vertical gradient of MSE between the two soil moisture regimes was greater in the fair weather case than in the convective case. If a preexisting disturbance were to force convection in the fair weather cases, the differences in precipitation would probably be larger between wet and dry regimes than in the convective case. Because fair weather soundings are typical of the HAPEX-Sahel region even for much of the wet season, further investigation of a perturbed fair weather environment is warranted.
Acknowledgments
The GCE model is supported by the NASA Physical Climate Program and the TRMM Project. We are grateful to Dr. R. Kakar (NASA HQ) for his support of GCE modeling research and to Pete Wetzel and Barry Lin for GCE-PLACE coupling. Data from the HAPEX Sahel Experiment were obtained from the Institut de Recherché pour le Développement (IRD). We thank Doug Parker, Rich Ellis, and Chris Taylor for providing a wealth of data. This work benefited greatly from discussions with Gareth Berry, Alan Srock, Michael Tanu, and Chris Thorncroft, as well as the comments of our three anonymous reviewers. The National Science Foundation supported our research under Grant NSF 0215413.
REFERENCES
Ambouta, K. J. M., 1984: Contribution à l′èdaphologie de la brousse tigrèe de l′Ouest nigèrien. Doctor-Engineer thesis, University of Nancy, 116 pp.
Betts, A. K., and Ball J. H. , 1995: The FIFE surface diurnal cycle climate. J. Geophys. Res., 100 , 25679–25693.
Betts, A. K., and Ball J. H. , 1998: FIFE surface climate and site-average dataset 1987–89. J. Atmos. Sci., 55 , 1091–1108.
Bromley, J., Brouwer J. , Barker A. P. , Gaze S. R. , and Valentin C. , 1997: The role of surface water redistribution in an area of patterned vegetation in a semi-arid environment, south-west Niger. J. Hydrol., 198 , 1–29.
Businger, J. A., Wyngaard J. C. , Izumi Y. , and Bradley E. F. , 1971: Flux-profile relationships in the atmospheric surface layer. J. Atmos. Sci., 28 , 181–189.
Cook, K. H., 1999: Generation of the African easterly jet and its role in determining West African precipitation. J. Climate, 12 , 1165–1184.
Eltahir, E. A. B., and Gong C. , 1996: Dynamics of wet and dry years in West Africa. J. Climate, 9 , 1030–1042.
Ferrier, B. S., Simpson J. , and Tao W-K. , 1996: Factors responsible for precipitation efficiencies in midlatitude and tropical squall simulations. Mon. Wea. Rev., 124 , 2100–2125.
Findell, K. L., and Eltahir E. A. B. , 2003: Atmospheric controls on soil moisture–boundary layer interactions. Part I: Framework development. J. Hydrometeor., 4 , 552–569.
Galle, S., Ehrmann M. , and Peugeot C. , 1999: Water balance in a banded vegetation pattern: A case study of tiger bush in western Niger. CATENA, 37 , 197–216.
Goutorbe, J-P., and Coauthors, 1994: HAPEX-Sahel: A large-scale study of land–atmosphere interactions in the semi-arid tropics. Ann. Geophys., 12 , 53–64.
Grabowski, W. W., Wu X. , Moncrieff M. W. , and Hall W. D. , 1998: Cloud-resolving modeling of cloud systems during phase III of GATE. Part II: Effects of resolution and the third spatial dimension. J. Atmos. Sci., 55 , 3264–3282.
Hodges, K. I., and Thorncroft C. D. , 1997: Distribution and statistics of African mesoscale convective weather systems based on the ISCCP Meteosat imagery. Mon. Wea. Rev., 125 , 2821–2837.
Klemp, J. B., and Wilhelmson R. , 1978: The simulation of three dimensional convective storm dynamics. J. Atmos. Sci., 35 , 1070–1096.
Lin, Y-L., Rarley R. D. , and Orville H. D. , 1983: Bulk parameterization of the snow field in a cloud model. J. Climate Appl. Meteor., 22 , 1065–1092.
Lucas, C., Zipser E. J. , and Ferrier B. S. , 2000: Sensitivity of tropical West Pacific oceanic convective lines to tropospheric wind and moisture profiles. J. Atmos. Sci., 57 , 2351–2373.
Lynn, B. H., Tao W-K. , and Wetzel P. J. , 1998: A study of landscape generated deep moist convection. Mon. Wea. Rev., 126 , 928–942.
Mohr, K. I., 2004: Interannual, monthly, and regional variability in the wet season diurnal cycle of precipitation in sub-Saharan Africa. J. Climate, 17 , 2441–2453.
Mohr, K. I., and Thorncroft C. , 2006: Intense convective systems in West Africa and their relationship to the African easterly jet. Quart. J. Roy. Meteor. Soc., 132 , 163–176.
Mohr, K. I., Baker R. D. , Tao W-K. , and Famiglietti J. S. , 2003: The sensitivity of West African convection line water budgets to land cover. J. Hydrometeor., 4 , 62–76.
Nicholls, M. E., Johnson R. H. , and Cotton W. R. , 1988: The sensitivity of two-dimensional simulations of tropical convective lines to environmental profiles. J. Atmos. Sci., 45 , 3625–3649.
Noilhan, J., and Planton S. , 1989: A simple parameterization of land surface processes for meteorological models. Mon. Wea. Rev., 117 , 536–549.
Rutledge, S. A., and Hobbs P. V. , 1984: The mesoscale and microscale structure and organization of clouds and precipitation in mid-latitude cyclones. Part XII: A diagnostic modeling study of precipitation development in narrow cold frontal rainbands. J. Atmos. Sci., 41 , 2949–2972.
Soong, S-T., and Ogura Y. , 1980: Response of trade wind cumula to large-scale processes. J. Atmos. Sci., 37 , 2035–2050.
Tao, W-K., and Simpson J. , 1993: Goddard Cumulus Ensemble Model. Part I: Model description. Terr. Atmos. Oceanic Sci., 4 , 35–72.
Tao, W-K., Lang S. , Simpson J. , Sui C-H. , Ferrier B. , and Chou M-D. , 1996: Mechanisms of cloud–radiation interaction in the Tropics and midlatitudes. J. Atmos. Sci., 53 , 2624–2651.
Tao, W-K., and Coauthors, 2003: Microphysics, radiation, and surface processes in the Goddard Cumulus Ensemble (GCE) model. Meteor. Atmos. Phys., 82 , 97–137.
Taylor, C. M., and Lebel T. , 1998: Observational evidence of persistent convective-scale rainfall patterns. Mon. Wea. Rev., 126 , 1597–1607.
Taylor, C. M., Harding R. J. , Thorpe A. J. , and Bessemoulin P. , 1997a: A mesoscale simulation of land-surface heterogeneity from HAPEX-Sahel. J. Hydrol., 188–189 , 1040–1066.
Taylor, C. M., Said F. , and Lebel T. , 1997b: Interactions between the land surface and meso-scale rainfall variability during HAPEX-Sahel. Mon. Wea. Rev., 125 , 2211–2227.
Taylor, C. M., Lambin E. F. , Stephenne N. , Harding R. J. , and Essery R. L. H. , 2002: The influence of land use change on climate in the Sahel. J. Climate, 15 , 3615–3629.
Thorncroft, C. D., and Blackburn M. , 1999: Maintenance of the African easterly jet. Quart. J. Roy. Meteor. Soc., 125 , 763–786.
Valentin, C., and d’Herbes J. M. , 1999: Niger tiger bush as a natural water harvesting system. CATENA, 37 , 231–256.
Wallace, J. S., and Holwill C. J. , 1997: Soil evaporation from tiger bush in south-west Niger. J. Hydrol., 188–189 , 426–442.
Wallace, J. S., and Coauthors, 1994: HAPEX-Sahel southern super-site report: An overview of the site and the experimental programme during the intensive observation period in 1992. Institute of Hydrology, Wallingford, United Kingdom, 55 pp.
Wetzel, P. J., and Boone A. , 1995: A Parameterization for Land–Atmosphere Cloud Exchange (PLACE): Documentation and testing of a detailed process model of the partly cloudy boundary layer over heterogeneous land. J. Climate, 8 , 1810–1837.
Wu, X. B., Thurow T. L. , and Whisenant S. G. , 2000: Fragmentation and changes in hydrologic function of tiger bush landscapes, south-west Niger. J. Ecol., 88 , 790–800.
Xu, K-M., and Randall D. A. , 1996: Explicit simulation of cumulus ensembles with the GATE Phase III data: Comparison with observations. J. Atmos. Sci., 53 , 3709–3736.
Xue, Y., and Shukla J. , 1993: The influence of land surface properties on Sahel climate. Part I: Desertification. J. Climate, 6 , 2232–2245.
Zilitinkevich, S. S., 1975: Comments on “A model for the dynamics of the inversion above a convective boundary layer.”. J. Atmos. Sci., 32 , 991–992.
Selected soil properties used in the numerical simulations after an offline spin up of the PLACE model. Note that initial soil moisture values are domain averages. Individual grid box soil moisture randomly varied from these mean values by ±0.02 cm3 cm−3.
Selected vegetation properties used in the numerical simulations. The % vegetation cover is the domain-average percentage of area covered by transpiring vegetation; root profile is the cumulative frequency distribution of roots in the five soil moisture layers. Vegetation cover randomly varied between 30% and 35% in this study to capture the variability in tiger bush environment.
Frequency distribution values (%) of binned surface precipitation rates.
Final water budgets for the wet and dry regimes within the 25 Aug 1992 case study. The units for C through Sc are 108 kg km−1.