1. Introduction
The land surface and atmosphere are linked through the exchange of water and energy, which modulates boundary layer development and precipitation that feed back to surface–atmosphere exchanges (e.g., Koster et al. 2004; Seneviratne et al. 2006). One example of land–atmosphere feedbacks with implications for land management, water resources, and regional climate is found in the North American northern Great Plains. Since the 1970s, shifts in cropping systems that include the reduction of summer fallow (Lubowski et al. 2006; Long et al. 2014; Vick et al. 2016) have coincided with a summertime cooling trend (Betts et al. 2013, 2014; Mahmood et al. 2014) of up to 2°C, a 7% increase in relative humidity (RH), and a 10 mm decade−1 increase in precipitation over parts of the Canadian Prairie provinces (Gameda et al. 2007) and moistening that extends to northeastern Montana and the Dakotas (Fig. 1).
The locations of the US-FPe eddy covariance research site and GGW atmospheric sounding site in Montana, with regional trends in specific humidity from 1970 until 2010 from the Climate Research Unit (CRU TS 3.10) database (Harris et al. 2014). Specific humidity was calculated using 2-m temperature, vapor pressure, and a 1000-hPa atmospheric reference pressure.
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
Such behavior is consistent with evidence for agriculture’s influence on weather and climate (reviewed in Raddatz 2007). Vegetated surfaces increase evapotranspiration [related via the latent heat of vaporization to latent heat flux (LE)] at the expense of sensible heat flux H. These changes in surface energy partitioning toward lower Bowen ratios (Bo = H/LE) result in shallower and moister atmospheric boundary layers (ABLs) compared to bare soils (e.g., Gameda et al. 2007; Vick et al. 2016) or in some cases natural vegetation (e.g., McPherson et al. 2004; Mahmood et al. 2014). Given that land management and land cover change have considerable effects on surface temperatures (Luyssaert et al. 2014; Mueller et al. 2015; Bright et al. 2017) and impact the water cycle through evapotranspiration and increases in atmospheric moisture, research is merited to further study local and regional effects of land–atmosphere interactions. In this context, Betts et al. (2013) noted that increased evapotranspiration in the Canadian Prairie provinces was responsible for a 0.34 mm day−1 increase in growing season precipitation. It remains unclear if changes in agricultural management that are similar to those in the Canadian Prairies have impacted precipitation processes in adjacent regions of the United States (Vick et al. 2016) and, if so, what mechanisms underlie these changes.
Generally speaking, coupling in the land–atmosphere system is well documented: the coevolution between temperature and moisture in the mixed layer as well as surface flux partitioning presents several direct feedbacks. Heating or drying of the ABL increases evaporative demand by intensifying the vapor pressure deficit (VPD), which in turn increases LE under well-watered conditions. The subsequent moistening of the ABL presents a negative feedback to evapotranspiration (van Heerwaarden et al. 2009, 2010). Similarly, vegetation responds to VPD through stomatal regulation of transpiration when VPD exceeds a threshold of approximately 10 hPa (Körner 1995; Oren et al. 1999; Lasslop et al. 2010). The partitioning of net radiation
Mixed layer (or slab type) models can provide valuable insight into these processes. They are computationally inexpensive and thus convenient for assessing land–atmosphere feedbacks and for quantifying the relationship between ecohydrological (e.g., surface flux partitioning) and atmospheric controls (e.g., atmospheric stability and moisture) on precipitation. ML models also have limitations. They are typically local in nature, such that regional circulation and large-scale effects are unaccounted for. Moreover, boundary layer structures and turbulent exchange across the capping inversion are prescribed, and accurate modeling of cloud-topped boundary layers proves challenging due to the latent heat release from condensation, convective mass flux, and radiative effects, all of which impact boundary layer dynamics and growth (e.g., Stull 1988). Despite these limitations, simplified ML models have proven to be a valuable tool to investigate interactions between convection and soil moisture (e.g., Ek and Holtslag 2004; Juang et al. 2007a,b; Siqueira et al. 2009; Konings et al. 2010, 2011), convection and the groundwater table (Bonetti et al. 2015), the regulation of convective cloud formation above forests (Manoli et al. 2016), and impacts of land management on h (Vick et al. 2016). However, ML models have also been applied to diagnose the surface exchange of heat from boundary layer characteristics (Santanello et al. 2005; Gentine et al. 2013b) and extended to include thermodynamic quantities such as convective available potential energy (CAPE) as additional diagnostics (Yin et al. 2015). Such enhancements further refine slab-type models with the caveat that tendencies in the atmospheric profile that occur due to advection or diabatic warming, both of which impact thermodynamics, are not considered. Similarly, Findell and Eltahir (2003a,b) developed a framework to assess the thermodynamic state and the potential for convective development from atmospheric profiles of temperature and moisture (CTP–
The present work combines a mixed layer modeling approach with the CTP–
2. Methodology
a. Data
Surface observations of H, LE,
Atmospheric sounding data were obtained for the period from 1975 to 2015 for the Glasgow International Airport in Montana (station code GGW), located approximately 110 km west of the Fort Peck site (48.21°N, 106.61°W). The airport is located to the northeast of the town of Glasgow and is surrounded by fields and grassland. Mixed layer evolution for the convective season, approximated here as the period between May and September, is based on the 1200 UTC profiles of T and moisture q, which correspond to approximately 0500 local time (LT). As ML development is driven by solar irradiance, all times used in this work are LT rather than UTC. Atmospheric lapse rates of potential temperature and humidity (
Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2; Rienecker et al. 2011; Reichle et al. 2017; Gelaro et al. 2017), data are used to asses climatic surface trends in the region. Specifically, MERRA-2 land surface diagnostics comprising LE, H, and
b. Model description


















Potential temperature and water vapor mixing ratio lapse rates [
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
The above equations constitute a closed system for modeling h,




c. Model setup
For each day, Bo is determined as the mean value between 1000 and 1500 LT, similar to Rigby et al. (2015), who suggested the use of the diurnal mean rather than daily maximum Bo to avoid overestimation of h. The maximum daytime net radiation
The chosen approach to model h constitutes a simple but tractable model, which excludes the radiative and thermodynamic effects of clouds (see, e.g., van Stratum et al. 2014), detailed atmospheric profiles, and dynamic processes such as advection and subsidence. However, the inclusion of these would greatly increase the degrees of freedom in the analysis, which are not constrained by observations.
d. Coupling metrics




3. Results
a. Observational results
Meteorological variables at the study site exhibited large seasonal cycles. From May through September, defined here as the convective season,
The observed median diurnal cycle of (a)
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
A look at quantiles of mean daytime Bo (Table 1) reveals that Bo can drop below 0.5 for a significant portion of days during June and July, whereas for the other months of the convective season Bo remained higher. At the upper end of the daytime Bo range, Bo values exceeding 2.75 made up more than 25% of days for all months except June.
Statistics of
The progressive summertime drying in northeastern Montana and the Fort Peck area are also evident in the precipitation data. Average precipitation at US-FPe decreased from more than 2 mm day−1 in May to less than 1 mm day−1 in September (not shown). Similarly the number of rain events also decreased.
As expected, θ increased with height (
b. Climatological trends
To characterize climatological trends in the Fort Peck region, we proceed to analyze MERRA-2 reanalysis data (Fig. 4). Monthly mean data show significant increasing trends for LE from 2.8 to 7.2 W m−2 decade−1 (at the
Climatic surface trends for May–September (left),(center) from MERRA-2 for the US-FPe and (right) from radiosondes at GGW for (a) H, (b) Bo, (c)
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
Linear trend and significance level (p value) for MERRA-2 surface variables at US-FPe for the period 1980–2015 and sounding lapse rates at GGW for the period 1973–2015.
c. Modeled h and LCL at US-FPe
During the convective seasons of 2000–08, there were only 63 days (out of 561 days passing quality control) with recorded daytime precipitation at US-FPe, highlighting the dry conditions in northeastern Montana. Of these, 19 occurrences were on days when modeled h exceeded the LCL. Overall, and despite the simplifications discussed above, the model appears to mostly capture the timing of locally developing precipitation events (Fig. 5a). The two events where precipitation occurred more than 5 h after h crossing the LCL were individually identified as large-scale precipitation events. Subsequently, the model is used to test whether the different model outcomes described in section 2b correspond to differences in environmental and atmospheric conditions, which is an important prerequisite for applying the model to study the sensitivity of the system to climatic and ecohydrological changes.
Summary of daily ML model results using US-FPe observations as model forcing: (a) timing of modeled ABL height h crossing the LCL (i.e., hxLCL) compared to the observed start of precipitation and grouping of boundary layer states (explained in text) based on environmental conditions such as (b) temperature and moisture lapse rates, (c) effective ML relative humidity, and (d) CTP–HIlow framework for groups: no observed precipitation and
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
Results suggest (Figs. 5b–d) that the convective outcomes of the model show little dependence on
Accepting that the model distinguishes between defined atmospheric and environmental states for locally developing convection, we frame our subsequent discussion in terms of the role of observed changes in regional hydrometeorological conditions on processes related to convective initiation.
4. Discussion
a. Observations
Observed seasonal patterns of LE and Bo (Fig. 3) can be used to characterize the ecohydrologic environment at US-FPe. From May to June, plant development increased evapotranspiration. Then, from June until August the environment became increasingly drier, due in part to plant senescence (Vick et al. 2016), thus reducing LE and increasing the average Bo to approximately 2.5 in September. High VPD can rapidly decrease Bo in grasslands, which can exhibit isohydric behavior (i.e., maintain near constant leaf water potential) and close stomatal control over transpiration (Novick et al. 2016; Konings et al. 2017). The results highlight the interactions between vegetation, environmental drivers, and ecophysiological responses to the land–atmosphere exchange of water, while the extremely high Bo values encountered during July–September demonstrate the frequently very dry surface conditions in the study area. The observed diurnal behavior of Bo is in reasonable agreement with the expectation of constant Bo during the daytime (Crago and Brutsaert 1996; Gentine et al. 2007, 2011), which is assumed by the ML model.
b. Climatological trends
Given that direct observations of surface energy balance and land–atmosphere exchange of energy and water are both labor intensive and expensive, their availability is inherently limited. To bridge this gap, this work applies MERRA-2 data, which are available from 1980 onward to assess climatic trends in the region (Figs. 4a–d). However, global gridded datasets such as MERRA-2 should be used with caution given that subgrid-scale surface heterogeneity, data assimilation, observational uncertainties, and the underlying modeling system cause local biases that affect water and energy cycles (e.g., Decker et al. 2012; Santanello et al. 2015). Nevertheless, in the absence of long-term observations, they can provide a first-order estimate of regional climatic trends, noting that the eddy-covariance site, which provides local measurements, might not be fully representative with respect to land cover, topography, or soil moisture of the MERRA-2 grid cell. Additionally, surface energy balance components in MERRA-2 are highly dependent on accurate prediction of local cloud cover, and MERRA-2 is demonstrated to have a positive bias in
Betts et al. (2013, 2014) reported increasing cloud cover in agricultural regions of the Canadian Prairie provinces, which is consistent with higher downward longwave radiation and lower solar irradiance. While we find a comparatively small effect on
c. Modeled h and LCL at US-FPe
The small portion of precipitating hxLCL events (Fig. 5) warrants discussion, as it might indicate overdetection of LCL crossings by the model. While misdetection of events cannot be fully excluded, note that summertime HIlow values are much higher than for previous studies (Findell and Eltahir 2003a,b), indicating that despite high observed H, the lack of low-level moisture in atmospheric profiles frequently controls convective development by suppressing the transition from shallow to precipitating convection, which is governed to a significant degree by the availability of moisture (Wu et al. 2009). For the sake of simplicity, the analytical model of Porporato (2009) is used in this work, which precludes the quantification of thermodynamic conditions under which the level of free convection is reached that require realistic profiles (Gerken et al. 2013) and can be assessed through integrated frameworks (see, e.g., Tawfik and Dirmeyer 2014; Tawfik et al. 2015). Also, eastward propagating mesoscale convective systems (MCSs) generated east of the Rocky Mountains (e.g., Tuttle and Davis 2006; Phillips and Klein 2014) are responsible for approximately 60% of total precipitation in the U.S. Great Plains (Carbone and Tuttle 2008). Because of their local nature, ML models, in general, cannot account for precipitation attributed to MCS, thus reducing the skill of the model to predict precipitation timing. This also limits direct feedbacks between local convection and soil moisture. However, because of the prevailing dryness in northeastern Montana, even moderate amounts of locally forced convective precipitation can be a crucial source of water for agriculture.
The ML model is used to assess environmental conditions governing the occurrence of hxLCL events and, furthermore, to establish whether certain environmental conditions are more likely to be associated with convective precipitation. Results demonstrate that both near-surface RH and Bo can be used to classify convective states at US-FPe. The fact that precipitation development for days with hxLCL is associated with smaller Bo values than for hxLCL events without precipitation is consistent with the notion that for small Bo (
d. Wet and dry coupling
The analysis of coupling regimes for GGW is broadly consistent with the results obtained by Findell and Eltahir (2003b). Figure 6a shows both wet and dry coupling behavior as indicated by the maximum of hxLCL probabilities for very low and very high Bo. At the same time, Fig. 6b reveals that the median state of the atmosphere is close to conditions where convection is suppressed mainly due to a lack of midlevel moisture, highlighting the role of atmospheric profiles in governing convective development near Fort Peck. Modeling results using the 75th percentile of
Model-estimated wet–dry coupling (defined in text) for the 75th percentile of
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
The coupling behavior and its relationship to atmospheric control are further illustrated with results from the CTP–HIlow framework (Fig. 6b). The median CTP–HIlow state observed at GGW is close to the boundary between atmospheric control preventing convective triggering and transitional states. Since 1975, median HIlow values have decreased by approximately 2°C, while trends in CTP were less clear (Table 3). As a consequence, coupling states moved from atmospheric control to a more transitional state. Note that despite using 40 years of sounding data (>6000 profiles), bins that correspond to CTP–HIlow combinations outside the interquartile range in Fig. 6b are sparsely populated or not populated by data, resulting in less clearly defined areas of coupling compared to Roundy et al. (2013), who used a much larger regional dataset. Nevertheless, compared to previous studies (Findell and Eltahir 2003b; Roundy et al. 2013), coupling behavior is found at more negative CTP values and higher HIlow values. Additionally, less defined areas of coupling also suggest that while the CTP–HIlow framework is useful to broadly characterize coupling, there are additional atmospheric factors that affect coupling behavior. Recent work by Cioni and Hohenegger (2017) showed that total precipitation amounts were always smaller during dry coupling compared to wet coupling, which is consistent with the notion that total column precipitable water rather than locally sourced moisture makes up the bulk of rainfall (e.g., Trenberth 1999). These issues suggest that modifications to the CTP–HIlow framework, which, given our focus on ML development is beyond the scope of this work, may help in better capturing convectively preconditioned states and resulting rainfall in northeastern Montana.
Statistics of observed CTP–HIlow from soundings at GGW.
5. Sensitivity of land–atmosphere coupling to climatic trends
While the underlying cause of the climatic trends affecting northeastern Montana as a whole are unclear, trends in Bo,
The comparison between the modeled h and LCL using sounding profiles for the decades from 1975 to 1985 and from 2005 to 2015 reveals a considerable increase in likelihood of hxLCL (Fig. 7) for a given level of
Modeled wet–dry coupling behavior as a function of
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
Trends in Bo also affect convective outcomes by increasing the probability of hxLCL and thus likely convection (for constant
There is also strong seasonal behavior in the environmental controls governing convectively preconditioned conditions as revealed by
Modeled sensitivity of hxLCL for (a)–(e) May–September and (f) the complete May–September convective season. The colored lines indicate the dividing line between clouds and no clouds for the range of Bo from 0.25 to 4.0. Initial conditions corresponding to
Citation: Journal of Hydrometeorology 19, 1; 10.1175/JHM-D-17-0117.1
In May during the beginning of the convective season, the boundary between modeled convective and nonconvective behavior is located within the interquartile range of atmospheric moisture characteristics. At the same time, the system shows a moderate to weak sensitivity with respect to Bo, where larger Bo requires higher
The notion that moisture limitation is a limiting factor in convection development is supported by the comparison between observed distributions of
This behavior agrees well with observed precipitation events in August and September at US-FPe, which are rare compared to rain events earlier in the season (results not shown) and the finding of reduced modeled total precipitation during dry coupling (Cioni and Hohenegger 2017). Last, the strong seasonality of behavior between May and September highlights the fact that the sensitivity between atmospheric moisture and energy partitioning should be examined on a subseasonal/monthly basis rather than for the convective season as a whole, since the season aggregated results (Fig. 8f) are greatly different from monthly results (Figs. 8a–e).
Our results suggest that convective precipitation occurs in northeastern Montana in response to atmospheric moistening and is becoming increasingly sensitive to Bo. Note the caveat that we assume convective precipitation, which we cannot assess directly, to behave like hxLCL following the notion that hxLCL is a “necessary but not sufficient condition” for convective initiation. The Bo tends to decrease throughout the convective season, whereas wet and dry coupling shifts from wet toward dry coupling in the late summer. In September, the absence of convection is governed by dry atmospheric profiles (atmospheric control). The corresponding hydrometeorological response to climate trends is thus likely increasing convective precipitation in the earlier growing season, but less in late summer.
While the results of one-dimensional models are useful to explore land–atmosphere coupling, it should be noted that only local effects are taken into account. Nonlocal precipitation events, for example, through eastward propagating MCSs (Phillips and Klein 2014; Carbone and Tuttle 2008; Tuttle and Davis 2006), cannot be addressed with this method. Also, it is well known that mesoscale circulations (e.g., forced from thermal or soil moisture differences) are associated with convective triggering over dry patches (e.g., Taylor et al. 2007; Garcia-Carreras et al. 2010), affecting land–atmosphere coupling as described by Koster et al. (2004), Seneviratne et al. (2010), and others. Similarly, cloud development impacts surface processes through cloud shading (e.g., Lohou and Patton 2014; Gronemeier et al. 2017) and boundary layer development through dynamic and radiative effects (e.g., Stull 1988). It is therefore desirable to merge local and regional methods, as done in Song et al. (2016), to investigate land–atmosphere coupling from the local to the regional scale, and to define the region of the North American northern Great Plains that has undergone regional climate responses that are consistent with shifts in agricultural management (Gameda et al. 2007; Betts et al. 2013, 2014; Mahmood et al. 2014).
6. Conclusions
This work applies a simplified analytical model of mixed layer heights and the lifting condensation level combined with the CTP–HIlow framework (Findell and Eltahir 2003a,b) to northeastern Montana, a region that has undergone considerable land cover change (Long et al. 2014). Motivated by documented climatic changes over the past four decades—namely, higher near-surface moisture amounts and increased partitioning of net radiation to latent heat fluxes at the expense of sensible heat as evidenced by smaller Bowen ratios—we examine how these trends affect coupling behavior.
Based on precipitation timing and the fact that the four convective outcomes [defined as 1) nonconvection permitting and no precipitation, 2) large-scale precipitation not controlled by local effects, 3) convection permitting without rain, and 4) convection permitting and precipitating] could be separated based on model initial conditions and the thermodynamic state of the atmospheres as characterized by the CTP–HIlow framework, the model is deemed to be useful to investigate the sensitivities of the system. However, CTP–HIlow alone cannot fully explain wet and dry coupling, suggesting that additional metrics are needed. Also, the very dry atmospheric conditions in August and September (suggesting atmospheric control on convection) and the small number of locally developed daytime precipitation events pose challenges to the modeling strategy yet emphasize the importance of additional convective season precipitation events to agricultural management. While mesoscale convective systems may be responsible for more than half of the total precipitation at Fort Peck, any additional precipitation from locally developing convection is likely to have a beneficial impact on crop yields.
Convectively preconditioned conditions near Fort Peck are closely associated with the availability of atmospheric moisture and sensible heat fluxes. Depending on tropospheric moisture contents and surface energy flux partitioning, mixed layer growth and associated entrainment of dry air can prevent the mixed layer height from reaching the LCL, while given adequate moisture supply, increased sensible heat fluxes are beneficial to reaching a convectively preconditioned state. As a consequence, the probability for convectively preconditioned conditions is smallest for intermediate Bowen ratios between approximately 0.5 and 2, indicating the presence of both wet and dry coupling. It is noteworthy that convectively preconditioned conditions occur at much drier conditions than proposed by Findell and Eltahir (2003a,b) and that the median state of the atmosphere is near the intersection point between moisture-limited suppressed convection as well as wet and dry coupling, highlighting the interplay between surface and atmospheric controls, which also exhibit seasonal dynamics. Over the course of the convective season the atmosphere transitions from wet coupling over dry coupling to atmospheric control, so that climatic trends suggest increased precipitation earlier in the season and less precipitation later on (August–September). At the same time, overall more convection is expected in response to regional moistening.
In the light of the climatic trend toward increased atmospheric moisture levels in the North American Great Plains (Pan et al. 2004) and the Great Plains’ importance to agricultural production, increased understanding of land–atmosphere coupling can help devise strategies for improved land management or climate adaptation. Future studies should quantify the area undergoing these changes in both surface and atmospheric dynamics and quantify how ongoing changes in agricultural management have and perhaps will continue to increase the likelihood of convective precipitation as approximated by mixed layer height and lifting condensation level crossings.
Acknowledgments
We acknowledge the Global Modeling and Assimilation Office (GMAO) and the GES DISC for the dissemination of MERRA. Funding for AmeriFlux data resources was provided by the U.S. Department of Energy’s Office of Science. CTP is calculated using the ctp_hi_low function made public by A.B. Tawfik (coupling-metrics.com; github.com/abtawfik/coupling-metrics.git). The University of Wyoming’s Atmospheric Science program and Larry Oolman are acknowledged for providing access to the radiosonde data (http://weather.uwyo.edu/upperair/sounding.html). We thank Tilden Meyer for eddy covariance data provision, and the AmeriFlux Management Project with the support of CDIAC for its harmonization. We acknowledge support from the National Science Foundation (NSF) Office of Integrated Activities (OIA) 1632810, the NSF Division of Environmental Biology (DEB) 1552976, the U.S. Department of Agriculture (USDA) National Institute of Food and Agriculture (NIFA) Hatch project 228396, the Montana Wheat and Barley Committee, and the graduate school at Montana State University. The authors thank Pierre Gentine and two anonymous reviewers for their helpful advice.
REFERENCES
Betts, A. K., R. Desjardins, D. Worth, and D. Cerkowniak, 2013: Impact of land use change on the diurnal cycle climate of the Canadian Prairies. J. Geophys. Res. Atmos., 118, 11 996–12 011, https://doi.org/10.1002/2013JD020717.
Betts, A. K., R. Desjardins, D. Worth, and B. Beckage, 2014: Climate coupling between temperature, humidity, precipitation, and cloud cover over the Canadian Prairies. J. Geophys. Res. Atmos., 119, 13 305–13 326, https://doi.org/10.1002/2014JD022511.
Bonetti, S., G. Manoli, J.-C. Domec, M. Putti, M. Marani, and G. G. Katul, 2015: The influence of water table depth and the free atmospheric state on convective rainfall predisposition: Water table and convective rainfall. Water Resour. Res., 51, 2283–2297, https://doi.org/10.1002/2014WR016431.
Bright, R. M., E. Davin, T. O’Halloran, J. Pongratz, K. Zhao, and A. Cescatti, 2017: Local temperature response to land cover and management change driven by non-radiative processes. Nat. Climate Change, 7, 296–302, https://doi.org/10.1038/nclimate3250.
Brimelow, J. C., J. M. Hanesiak, and W. R. Burrows, 2011: Impacts of land atmosphere feedbacks on deep, moist convection on the Canadian Prairies. Earth Interact., 15, https://doi.org/10.1175/2011EI407.1.
Carbone, R. E., and J. D. Tuttle, 2008: Rainfall occurrence in the U.S. warm season: The diurnal cycle. J. Climate, 21, 4132–4146, https://doi.org/10.1175/2008JCLI2275.1.
Cioni, G., and C. Hohenegger, 2017: Effect of soil moisture on diurnal convection and precipitation in large-eddy simulations. J. Hydrometeor., 18, 1885–1903, https://doi.org/10.1175/JHM-D-16-0241.1.
Crago, R., and W. Brutsaert, 1996: Daytime evaporation and the self-preservation of the evaporative fraction and the Bowen ratio. J. Hydrol., 178, 241–255, https://doi.org/10.1016/0022-1694(95)02803-X.
Davin, E. L., S. I. Seneviratne, P. Ciais, A. Olioso, and T. Wang, 2014: Preferential cooling of hot extremes from cropland albedo management. Proc. Natl. Acad. Sci. USA, 111, 9757–9761, https://doi.org/10.1073/pnas.1317323111.
Decker, M., M. A. Brunke, Z. Wang, K. Sakaguchi, X. Zeng, and M. G. Bosilovich, 2012: Evaluation of the reanalysis products from GSFC, NCEP, and ECMWF using flux tower observations. J. Climate, 25, 1916–1944, https://doi.org/10.1175/JCLI-D-11-00004.1.
Dirmeyer, P. A., Z. Wang, M. J. Mbuh, and H. E. Norton, 2014: Intensified land surface control on boundary layer growth in a changing climate. Geophys. Res. Lett., 41, 1290–1294, https://doi.org/10.1002/2013GL058826.
Draper, C. S., R. H. Reichle, and R. D. Koster, 2018: Assessment of MERRA-2 land surface energy flux estimates. J. Climate, 31, 671–691, https://doi.org/10.1175/JCLI-D-17-0121.1.
Driedonks, A. G. M., 1982: Models and observations of the growth of the atmospheric boundary layer. Bound.-Layer Meteor., 23, 283–306, https://doi.org/10.1007/BF00121117.
Ek, M. B., and A. M. Holtslag, 2004: Influence of soil moisture on boundary layer cloud development. J. Hydrometeor., 5, 86–99, https://doi.org/10.1175/1525-7541(2004)005<0086:IOSMOB>2.0.CO;2.
Ferguson, C. R., and E. F. Wood, 2011: Observed land–atmosphere coupling from satellite remote sensing and reanalysis. J. Hydrometeor., 12, 1221–1254, https://doi.org/10.1175/2011JHM1380.1.
Ferguson, C. R., and G. Villarini, 2012: Detecting inhomogeneities in the Twentieth Century Reanalysis over the central United States. J. Geophys. Res., 117, D05123, https://doi.org/10.1029/2011JD016988.
Findell, K. L., and E. A. B. Eltahir, 2003a: Atmospheric controls on soil moisture-boundary layer interactions. Part I: Framework development. J. Hydrometeor., 4, 552–569, https://doi.org/10.1175/1525-7541(2003)004<0552:ACOSML>2.0.CO;2.
Findell, K. L., and E. A. B. Eltahir, 2003b: Atmospheric controls on soil moisture-boundary layer interactions. Part II: Feedbacks within the continental United States. J. Hydrometeor., 4, 570–583, https://doi.org/10.1175/1525-7541(2003)004<0570:ACOSML>2.0.CO;2.
Ford, T. W., A. D. Rapp, and S. M. Quiring, 2015: Does afternoon precipitation occur preferentially over dry or wet soils in Oklahoma? J. Hydrometeor., 16, 874–888, https://doi.org/10.1175/JHM-D-14-0005.1.
Gameda, S., B. Qian, C. Campbell, and R. Desjardins, 2007: Climatic trends associated with summerfallow in the Canadian Prairies. Agric. For. Meteor., 142, 170–185, https://doi.org/10.1016/j.agrformet.2006.03.026.
Garcia-Carreras, L., D. J. Parker, C. M. Taylor, C. E. Reeves, and J. G. Murphy, 2010: Impact of mesoscale vegetation heterogeneities on the dynamical and thermodynamic properties of the planetary boundary layer. J. Geophys. Res., 115, D03102, https://doi.org/10.1029/2009JD012811.
Gelaro, R., and Coauthors, 2017: The Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2). J. Climate, 30, 5419–5454, https://doi.org/10.1175/JCLI-D-16-0758.1.
Gentine, P., D. Entekhabi, A. Chehbouni, G. Boulet, and B. Duchemin, 2007: Analysis of evaporative fraction diurnal behaviour. Agric. For. Meteor., 143, 13–29, https://doi.org/10.1016/j.agrformet.2006.11.002.
Gentine, P., D. Entekhabi, and J. Polcher, 2011: The diurnal behavior of evaporative fraction in the soil–vegetation–atmospheric boundary layer continuum. J. Hydrometeor., 12, 1530–1546, https://doi.org/10.1175/2011JHM1261.1.
Gentine, P., C. R. Ferguson, and A. A. M. Holtslag, 2013a: Diagnosing evaporative fraction over land from boundary-layer clouds. J. Geophys. Res. Atmos., 118, 8185–8196, https://doi.org/10.1002/jgrd.50416.
Gentine, P., A. A. M. Holtslag, F. D’Andrea, and M. Ek, 2013b: Surface and atmospheric controls on the onset of moist convection over land. J. Hydrometeor., 14, 1443–1462, https://doi.org/10.1175/JHM-D-12-0137.1.
Gerken, T., W. Babel, F. Sun, M. Herzog, Y. Ma, T. Foken, and H.-F. Graf, 2013: Uncertainty in atmospheric profiles and its impact on modeled convection development at Nam Co Lake, Tibetan Plateau. J. Geophys. Res. Atmos., 118, 12 317–12 331, https://doi.org/10.1002/2013JD020647.
Gronemeier, T., F. Kanani-Shring, and S. Raasch, 2017: Do shallow cumulus clouds have the potential to trigger secondary circulations via shading? Bound.-Layer Meteor., 162, 143–169, https://doi.org/10.1007/s10546-016-0180-7.
Guillod, B. P., B. Orlowsky, D. G. Miralles, A. J. Teuling, and S. I. Seneviratne, 2015: Reconciling spatial and temporal soil moisture effects on afternoon rainfall. Nat. Commun., 6, 6443, https://doi.org/10.1038/ncomms7443.
Harris, I., P. Jones, T. Osborn, and D. Lister, 2014: Updated high-resolution grids of monthly climatic observations the CRU TS3.10 dataset. Int. J. Climatol., 34, 623–642, https://doi.org/10.1002/joc.3711.
Juang, J.-Y., G. G. Katul, A. Porporato, P. C. Stoy, M. S. Siqueira, M. Detto, H.-S. Kim, and R. Oren, 2007a: Eco-hydrological controls on summertime convective rainfall triggers. Global Change Biol., 13, 887–896, https://doi.org/10.1111/j.1365-2486.2007.01315.x.
Juang, J.-Y., A. Porporato, P. C. Stoy, M. S. Siqueira, A. C. Oishi, M. Detto, H.-S. Kim, and G. G. Katul, 2007b: Hydrologic and atmospheric controls on initiation of convective precipitation events. Water Resour. Res., 43, W03421, https://doi.org/10.1029/2006WR004954.
Konings, A. G., G. G. Katul, and A. Porporato, 2010: The rainfall-no rainfall transition in a coupled land-convective atmosphere system. Geophys. Res. Lett., 37, L14401, https://doi.org/10.1029/2010GL043967.
Konings, A. G., S. C. Dekker, M. Rietkerk, and G. G. Katul, 2011: Drought sensitivity of patterned vegetation determined by rainfall-land surface feedbacks. J. Geophys. Res., 116, G04008, https://doi.org/10.1029/2011JG001748.
Konings, A. G., A. P. Williams, and P. Gentine, 2017: Sensitivity of grassland productivity to aridity controlled by stomatal and xylem regulation. Nat. Geosci., 10, 284–288, https://doi.org/10.1038/ngeo2903.
Körner, C., 1995: Leaf diffusive conductances in the major vegetation types of the globe. Ecophysiology of Photosynthesis, E.-D. Schulze and M. M. Caldwell, Eds., Springer Study Edition, Vol. 100, Springer, 463–490, https://doi.org/10.1007/978-3-642-79354-7_22.
Koster, R. D., and Coauthors, 2004: Regions of strong coupling between soil moisture and precipitation. Science, 305, 1138–1140, https://doi.org/10.1126/science.1100217.
Kustas, W. P., T. J. Schmugge, K. S. Humes, T. J. Jackson, R. Parry, M. A. Weltz, and M. S. Moran, 1993: Relationships between evaporative fraction and remotely sensed vegetation index and microwave brightness temperature for semiarid rangelands. J. Appl. Meteor., 32, 1781–1790, https://doi.org/10.1175/1520-0450(1993)032<1781:RBEFAR>2.0.CO;2.
Lasslop, G., M. Reichstein, D. Papale, A. D. Richardson, A. Arneth, A. Barr, P. Stoy, and G. Wohlfahrt, 2010: Separation of net ecosystem exchange into assimilation and respiration using a light response curve approach: Critical issues and global evaluation. Global Change Biol., 16, 187–208, https://doi.org/10.1111/j.1365-2486.2009.02041.x.
Lohou, F., and E. G. Patton, 2014: Surface energy balance and buoyancy response to shallow cumulus shading. J. Atmos. Sci., 71, 665–682, https://doi.org/10.1175/JAS-D-13-0145.1.
Long, J. A., R. L. Lawrence, P. R. Miller, L. A. Marshall, and M. C. Greenwood, 2014: Adoption of cropping sequences in northeast Montana: A spatio-temporal analysis. Agric. Ecosyst. Environ., 197, 77–87, https://doi.org/10.1016/j.agee.2014.07.022.
Lubowski, R. N., M. Vesterby, S. Bucholtz, A. Baez, and M. J. Robterts, 2006: Major uses of land in the United States, 2002. USDA Economic Information Bulletin 14, 47 pp., https://www.ers.usda.gov/webdocs/publications/43967/13011_eib14_1_.pdf?v=42061.
Luyssaert, S., and Coauthors, 2014: Land management and land-cover change have impacts of similar magnitude on surface temperature. Nat. Climate Change, 4, 389–393, https://doi.org/10.1038/nclimate2196.
Lytinska, Z., J. Parfiniewicz, and H. Piwkowski, 1976: The prediction of air mass thunderstorms and hails. Proc. WMO Symp. on the Interpretation of Broad-Scale NWP Product for Local Forecasting Purposes, WMO/TD-450, Warsaw, Poland, WMO, 128–130.
Mahmood, R., and Coauthors, 2014: Land cover changes and their biogeophysical effects on climate. Int. J. Climatol., 34, 929–953, https://doi.org/10.1002/joc.3736.
Manoli, G., J.-C. Domec, K. Novick, A. C. Oishi, A. Noormets, M. Marani, and G. Katul, 2016: Soil–plant–atmosphere conditions regulating convective cloud formation above southeastern US pine plantations. Global Change Biol., 22, 2238–2254, https://doi.org/10.1111/gcb.13221.
Marquardt Collow, A. B., and M. A. Miller, 2016: The seasonal cycle of the radiation budget and cloud radiative effect in the Amazon rain forest of Brazil. J. Climate, 29, 7703–7722, https://doi.org/10.1175/JCLI-D-16-0089.1.
McPherson, R. A., D. J. Stensrud, and K. C. Crawford, 2004: The impact of Oklahoma’s winter wheat belt on the mesoscale environment. Mon. Wea. Rev., 132, 405–421, https://doi.org/10.1175/1520-0493(2004)132<0405:TIOOWW>2.0.CO;2.
Meyers, T., 2000: AmeriFlux US-FPe Fort Peck. AmeriFlux, accessed 23 September 2016, https://doi.org/10.17190/AMF/1246053.
Mueller, N. D., E. E. Butler, K. A. McKinnon, A. Rhines, M. Tingley, N. M. Holbrook, and P. Huybers, 2015: Cooling of US Midwest summer temperature extremes from cropland intensification. Nat. Climate Change, 6, 317–322, https://doi.org/10.1038/nclimate2825.
Novick, K. A., and Coauthors, 2016: The increasing importance of atmospheric demand for ecosystem water and carbon fluxes. Nat. Climate Change, 6, 1023–1027, https://doi.org/10.1038/nclimate3114.
Oren, R., J. S. Sperry, G. G. Katul, D. E. Pataki, B. E. Ewers, N. Phillips, and K. V. R. Schäfer, 1999: Survey and synthesis of intra- and interspecific variation in stomatal sensitivity to vapour pressure deficit. Plant Cell Environ., 22, 1515–1526, https://doi.org/10.1046/j.1365-3040.1999.00513.x.
Pan, Z., R. W. Arritt, E. S. Takle, W. J. Gutowski, C. J. Anderson, and M. Segal, 2004: Altered hydrologic feedback in a warming climate introduces a warming hole. Geophys. Res. Lett., 31, L17109, https://doi.org/10.1029/2004GL020528.
Phillips, T. J., and S. A. Klein, 2014: Land-atmosphere coupling manifested in warm-season observations on the U.S. Southern Great Plains. J. Geophys. Res. Atmos., 119, 509–528, https://doi.org/10.1002/2013JD020492.
Porporato, A., 2009: Atmospheric boundary-layer dynamics with constant Bowen ratio. Bound.-Layer Meteor., 132, 227–240, https://doi.org/10.1007/s10546-009-9400-8.
Raddatz, R. L., 1993: Prairie agroclimate boundary-layer model: A simulation of the atmosphere/crop-soil interface. Atmos.–Ocean, 31, 399–419, https://doi.org/10.1080/07055900.1993.9649478.
Raddatz, R. L., 2007: Evidence for the influence of agriculture on weather and climate through the transformation and management of vegetation: Illustrated by examples from the Canadian Prairies. Agric. For. Meteor., 142, 186–202, https://doi.org/10.1016/j.agrformet.2006.08.022.
Reichle, R. H., C. S. Draper, Q. Liu, M. Girotto, S. P. P. Mahanama, R. D. Koster, and G. J. M. De Lannoy, 2017: Assessment of MERRA-2 land surface hydrology estimates. J. Climate, 30, 2937–2960, https://doi.org/10.1175/JCLI-D-16-0720.1.
Rienecker, M. M., and Coauthors, 2011: MERRA: NASAs Modern-Era Retrospective Analysis for Research and Applications. J. Climate, 24, 3624–3648, https://doi.org/10.1175/JCLI-D-11-00015.1.
Rigby, J. R., J. Yin, J. D. Albertson, and A. Porporato, 2015: Approximate analytical solution to diurnal atmospheric boundary-layer growth under well-watered conditions. Bound.-Layer Meteor., 156, 73–89, https://doi.org/10.1007/s10546-015-0018-8.
Roundy, J. K., C. R. Ferguson, and E. F. Wood, 2013: Temporal variability of land–atmosphere coupling and its implications for drought over the southeast United States. J. Hydrometeor., 14, 622–635, https://doi.org/10.1175/JHM-D-12-090.1.
Roundy, J. K., C. R. Ferguson, and E. F. Wood, 2014: Impact of land-atmospheric coupling in CFSv2 on drought prediction. Climate Dyn., 43, 421–434, https://doi.org/10.1007/s00382-013-1982-7.
Santanello, J. A., M. A. Friedl, and W. P. Kustas, 2005: An empirical investigation of convective planetary boundary layer evolution and its relationship with the land surface. J. Appl. Meteor., 44, 917–932, https://doi.org/10.1175/JAM2240.1.
Santanello, J. A., J. Roundy, and P. A. Dirmeyer, 2015: Quantifying the land-atmosphere coupling behavior in modern reanalysis products over the U.S. Southern Great Plains. J. Climate, 28, 5813–5829, https://doi.org/10.1175/JCLI-D-14-00680.1.
Segal, M., R. W. Arritt, C. Clark, R. Rabin, and J. Brown, 1995: Scaling evaluation of the effect of surface characteristics on potential for deep convection over uniform terrain. Mon. Wea. Rev., 123, 383–400, https://doi.org/10.1175/1520-0493(1995)123<0383:SEOTEO>2.0.CO;2.
Seneviratne, S. I., D. Lüthi, M. Litschi, and C. Schär, 2006: Land atmosphere coupling and climate change in Europe. Nature, 443, 205–209, https://doi.org/10.1038/nature05095.
Seneviratne, S. I., T. Corti, E. L. Davin, M. Hirschi, E. B. Jaeger, I. Lehner, B. Orlowsky, and A. J. Teuling, 2010: Investigating soil moisture climate interactions in a changing climate: A review. Earth-Sci. Rev., 99, 125–161, https://doi.org/10.1016/j.earscirev.2010.02.004.
Siqueira, M., G. Katul, and A. Porporato, 2009: Soil moisture feedbacks on convection triggers: The role of soil–plant hydrodynamics. J. Hydrometeor., 10, 96–112, https://doi.org/10.1175/2008JHM1027.1.
Song, H.-J., C. R. Ferguson, and J. K. Roundy, 2016: Land–atmosphere coupling at the Southern Great Plains Atmospheric Radiation Measurement (ARM) field site and its role in anomalous afternoon peak precipitation. J. Hydrometeor., 17, 541–556, https://doi.org/10.1175/JHM-D-15-0045.1.
Stull, R., 1988: An Introduction to Boundary Layer Meteorology. Springer, 670 pp.
Tawfik, A. B., and P. A. Dirmeyer, 2014: A process-based framework for quantifying the atmospheric preconditioning of surface-triggered convection. Geophys. Res. Lett., 41, 173–178, https://doi.org/10.1002/2013GL057984.
Tawfik, A. B., P. A. Dirmeyer, and J. A. Santanello, 2015: The heated condensation framework. Part II: Climatological behavior of convective initiation and land–atmosphere coupling over the conterminous United States. J. Hydrometeor., 16, 1946–1961, https://doi.org/10.1175/JHM-D-14-0118.1.
Taylor, C. M., D. J. Parker, and P. P. Harris, 2007: An observational case study of mesoscale atmospheric circulations induced by soil moisture. Geophys. Res. Lett., 34, L15801, https://doi.org/10.1029/2007GL030572.
Tennekes, H., 1973: A model for the dynamics of the inversion above a convective boundary layer. J. Atmos. Sci., 30, 558–567, https://doi.org/10.1175/1520-0469(1973)030<0558:AMFTDO>2.0.CO;2.
Trenberth, K. E., 1999: Atmospheric moisture recycling: Role of advection and local evaporation. J. Climate, 12, 1368–1381, https://doi.org/10.1175/1520-0442(1999)012<1368:AMRROA>2.0.CO;2.
Tuttle, J. D., and C. A. Davis, 2006: Corridors of warm season precipitation in the central United States. Mon. Wea. Rev., 134, 2297–2317, https://doi.org/10.1175/MWR3188.1.
van Heerwaarden, C. C., J. Vilà-Guerau de Arellano, A. F. Moene, and A. A. M. Holtslag, 2009: Interactions between dry-air entrainment, surface evaporation and convective boundary-layer development. Quart. J. Roy. Meteor. Soc., 135, 1277–1291, https://doi.org/10.1002/qj.431.
van Heerwaarden, C. C., J. Vilà-Guerau de Arellano, A. Gounou, F. Guichard, and F. Couvreux, 2010: Understanding the daily cycle of evapotranspiration: A method to quantify the influence of forcings and feedbacks. J. Hydrometeor., 11, 1405–1422, https://doi.org/10.1175/2010JHM1272.1.
van Stratum, B. J. H., J. Vilà-Guerau de Arellano, C. C. van Heerwaarden, and H. G. Ouwersloot, 2014: Subcloud-layer feedbacks driven by the mass flux of shallow cumulus convection over land. J. Atmos. Sci., 71, 881–895, https://doi.org/10.1175/JAS-D-13-0192.1.
Vick, E. S., P. C. Stoy, A. C. Tang, and T. Gerken, 2016: The surface-atmosphere exchange of carbon dioxide, water, and sensible heat across a dryland wheat-fallow rotation. Agric. Ecosyst. Environ., 232, 129–140, https://doi.org/10.1016/j.agee.2016.07.018.
Wu, C.-M., B. Stevens, and A. Arakawa, 2009: What controls the transition from shallow to deep convection? J. Atmos. Sci., 66, 1793–1806, https://doi.org/10.1175/2008JAS2945.1.
Yamada, H., 2008: Numerical simulations of the role of land surface conditions in the evolution and structure of summertime thunderstorms over a flat highland. Mon. Wea. Rev., 136, 173–188, https://doi.org/10.1175/2007MWR2053.1.
Yin, J., J. D. Albertson, J. R. Rigby, and A. Porporato, 2015: Land and atmospheric controls on initiation and intensity of moist convection: CAPE dynamics and LCL crossings. Water Resour. Res., 51, 8476–8493, https://doi.org/10.1002/2015WR017286.