Zum Hauptinhalt springen

Projection of vegetation impacts on future droughts over West Africa using a coupled RegCM-CLM-CN-DV

Um, Myoung-Jin ; Erfanian, Amir ; et al.
In: Climatic Change, Jg. 163 (2020-11-01), S. 653-668
Online unknown

Projection of vegetation impacts on future droughts over West Africa using a coupled RegCM-CLM-CN-DV 

This study investigates the projected effect of vegetation feedback on drought conditions in West Africa using a regional climate model coupled with the National Center for Atmospheric Research Community Land Model, carbon–nitrogen (CN) module, and dynamic vegetation (DV) module (RegCM4.3.4-CLM-CN-DV). The role of vegetation feedback is examined based on simulations with and without the DV module. Simulations from four different global climate models are used as lateral boundary conditions for historical and future periods (historical: 1981–2000; future: 2081–2100). Using the standardized precipitation evapotranspiration index (SPEI), we quantify the frequency, duration, and intensity of droughts over two focal regions of the Sahel and the Guinea Coast in West Africa. With the dynamic vegetation (DV) considered, future droughts are expected to become more prolonged and enhanced over the Sahel. The drought is defined when the SPEI value is lower than the threshold (− 1) in the simulations with DV than those without DV. Over the Guinea Coast, the impact of DV is opposite compared with that over the Sahel. Additionally, we show that the simulated annual leaf area index associates well with annual minimum SPEI with the correlation coefficient up to 0.91, particularly over the Sahel, which is a transition zone, where land–atmosphere feedback is relatively strong. Results signify the importance of vegetation dynamics in predicting future droughts in West Africa, where the biosphere and atmosphere interactions play a significant role in the regional climate regime.

Keywords: West Africa; SPEI; LAI; Vegetation feedback; Drought

Supplementary Information The online version of this article (https://doi.org/10.1007/s10584-020-02879-z) contains supplementary material, which is available to authorized users.

Introduction

West Africa is extremely vulnerable to climate change and projecting its future climate is a challenging task (Cook et al. [14]). From the 1970s, a long period of drought was observed over West Africa, lasting until the late 1990s. There is no clear consensus about whether the future of the West African hydroclimate will be drier or wetter. Some studies projected drying trends (Hulme et al. [30]), whereas others predicted a wetter future (Hoerling et al. [27]; Kamga et al. [32]; Maynard et al. [41]). Caminade and Terray ([9]) examined the simulated rainfall over the Sahel at the end of twenty-first century with 21 models from the Coupled Model Intercomparison Project (CMIP) Phase 3 (CMIP3). They argued that different model projections are highly uncertain because future rainfall may be affected by changes in surface conditions (e.g., vegetation, land use, and soil moisture) that have not been considered in CMIP3 models. Roehrig et al. ([54]) combined the CMIP3 and CMIP Phase 5 (CMIP5) global climate models (GCM) and found that the western Sahel shows a drying trend, whereas the eastern Sahel shows an opposite trend.

Limited-area models, i.e., regional climate models (RCMs), which are forced with lateral boundary conditions (LBCs) derived from GCMs, are often used as they can capture finer-resolution details, compared with GCMs (Kumar et al. [35]) since the physics of RCMs dominate the signals imposed by large-scale forcings (i.e., LBCs from GCMs). Akinsanola and Zhou ([2]) investigated projected changes in extreme summer rainfall events over West Africa with data from the Coordinated Regional Climate Downscaling Experiment (CORDEX) models. Results showed the RCMs participating in CORDEX reasonably reproduced the observed pattern of extreme rainfall over the region. Future projections under the representative concentration pathways showed a statistically significant decrease in total rainfall and an increase in the consecutive dry days and extreme rainfall. In another study, Akinsanola and Zhou ([3]) projected future changes in summer monsoon rainfall under the RCP8.5 scenario using the Rossby Centre Regional Climate Model (RCA4), driven by six CMIP5 GCM scenarios. The authors found that summer monsoon rainfall in a future period (2074–2099) is projected to increase than that in present-day period (1980–2005) over most parts of West Africa. In addition, Dixon et al. ([19]) studied the changes in monsoon precipitation over the Sahel region of West Africa under a global sea surface temperature (SST) increase by 4 K using a 2D-model by Peyrillé et al. ([53]). The results indicated that direct forcing would shift the precipitation band toward the south, whereas forcing associated with a global increase in SST would reduce precipitation.

Because climate and greenhouse gas concentrations continuously change, changes are expected in the vegetation as well (Yu et al. [74]). Thus, global and regional climate models should incorporate more representative and reliable prognostic vegetation dynamics instead of prescribing vegetation composition and structure, particularly for the regions where biosphere–atmosphere interactions are important (Alo and Wang [4]; Patricola and Cook [51]; Wramneby et al. [70]; Xue et al. [72]; Zhang et al. [79]). Charney et al. ([10]) first suggested that precipitation could change dynamically in response to vegetation changes, and suggested that the changes in precipitation over the Sahel were due to reduction in vegetation and increase in albedo. Various studies documented biosphere–atmosphere interactions (Kim and Wang [34]; Patricola and Cook [50]; Wang and Eltahir [66]), but relatively few RCM studies considered prognostic vegetation dynamics, (Cook and Vizy [13]; Garnaud et al. [22]; Wang et al. [67]; Yu et al. [75]).

Cook and Vizy ([13]) developed a vegetation model coupled with an RCM to estimate the influence of global warming on South America by allowing interactions between climate and vegetation. With the simulation of the future climate under the CMIP3 A2 scenario, the authors found a reduction in vegetation cover of approximately 70% in the Amazon rainforest along with a widespread increase in grass and shrubland in the east by the end of twenty-first century. This highlights the importance of considering vegetation dynamics in RCMs. Garnaud et al. ([22]) combined the Canadian Regional Climate Model (CRCM5) with the Canadian Territorial Ecosystem Model to investigate the impact of a vegetation model for simulating the present-day climate over North America. The results showed that including vegetation dynamics improved the model's performance in some regions while introducing new biases in other regions, owing to biases in the simulated leaf area index (LAI, a dimensionless quantity defined as the green leaf area per unit ground surface area). This atmosphere-vegetation interaction also introduced long-term memory, which was estimated using a lagged correlation between temperature/precipitation and LAI. Wu et al. ([71]) utilized RCA-GUESS (Smith et al. [56]), which is a regional earth system model coupled with a dynamic vegetation model, and investigated the role of vegetation dynamics in climate of Africa under the RCP8.5 projected climate scenario. The authors showed that including vegetation processes amplified the warming trend and enhanced precipitation reduction over rainforest areas, which highlights the impact of including vegetation processes in a climate model.

Recently, Wang et al. ([67]) introduced dynamic vegitation (DV) into the International Center for Theoretical Physics Regional Climate Model (RegCM4.3.4, Giorgi et al. [24]) with carbon–nitrogen (CN) and DV schemes (RegCM-CLM-CN-DV) of the community land model (CLM4.5, Lawrence et al. [36]; Oleson et al. [46]), and validated the coupled model over tropical Africa. With the RegCM-CLM-CN-DV, Yu et al. ([75]) and Erfanian et al. ([21]) examined the impacts of vegetation dynamics on the climate using multiple LBCs from past and future GCM simulations over West Africa. Yu et al. ([75]) showed that climate impact of dynamic vegetation feedback was found mainly in semiarid areas of West Africa with little signal in the wet tropics. Similarly, Erfanian et al. ([21]) also demonstrated substantial sensitivity of the simulated precipitation, evapotranspiration, and soil moisture to vegetation representation. Including DV in the model eliminates potential inconsistencies between prescribed vegetation and climate, but it can cause climate drift and enhance model biases (Erfanian et al. [21]).

While many studies focused on projecting future climate over West Africa, a limited number of studies quantitatively assessed drought frequency, duration, and intensity, as well as considering the role of vegetation dynamics on the drought processes in this region. In view of the uncertain future of drought-prone regions, this study aims to understand the impacts of vegetation feedback on the future of droughts over West Africa. Specifically, the standardized precipitation evapotranspiration index (SPEI) is used to depict the vegetation feedbacks on drought characteristics according to frequency, intensity, and duration over West Africa. Following previous studies with RegCM-CLM-CN-DV (Wang et al. [67]; Yu et al. [75]; Erfanian et al. [21]), we examined the drought characteristics simulated with and without vegetation dynamics for the historical and future periods and showed the signals of DV on the drought characteristics in two selected regions of West Africa.

Methodology

Model description

This study uses state-of-the-art RegCM-CLM-CN-DV (Wang et al. [67]). Specifically, RegCM4.3.4 (Giorgi et al. [24]) and CLM4.5 (Lawrence et al. [36]; Oleson et al. [46]) with CN dynamics and DV are coupled to simulate various atmospheric, land, biogeochemical, vegetation phenology, and vegetation distribution processes. RegCM is a regional model that uses an Arakawa B-grid finite differencing algorithm along with a terrain-following σ-pressure vertical coordinate system. The model uses the Grell et al. ([26]) dynamics from the hydrostatic version of the Pennsylvania State University Mesoscale Model version 5, and the radiation scheme from the Community Climate Model (Kiehl et al. [33]). The model includes four different convection parameterization schemes, namely (1) the modified-Kuo scheme (Anthes et al. [5]), (2) Tiedtke scheme (Tiedtke [62]), (3) Grell scheme (Grell [25]), and (4) Emanuel scheme (Emanuel [20]) along with the non-local boundary layer scheme of Holtslag et al. ([28]). The cloud and precipitation schemes come from Pal et al. ([48]). The aerosols algorithm follows Solmon et al. ([58]) and Zakey et al. ([76]).

To solve the various land processes in the model (e.g., surface biogeochemical and biogeophysical processes, ecosystem dynamics, and hydrological processes), CLM4.5 considers 15 soil layers, 16 distinct plant functional types (PFT), and up to 5 snow layers in each grid cell (Lawrence et al. [36]). Optional components available in this model are the CN and DV modules. The CN module not only simulates CN cycles and plant phenology but also estimates vegetation height, stem area index, and LAI. The DV module projects the fractional coverage of different PFTs and corresponding temporal variations at yearly time steps based on carbon budget from CN. If CN and DV modules are inactive, the vegetation distribution and composition in the model are prescribed according to the observed data sets (i.e., static vegetation, hereafter referred to as SV).

Numerical experiments

A total of 18 different numerical simulations are used in this study, as listed in Table 1, with 2 evaluation runs and 16 experimental runs with the climate change scenarios. Numerical simulations are conducted in two distinct configurations, one in which the CN-DV module is activated (i.e., DV runs) and the other in which the CN-DV module is not activated (i.e., SV runs). Additionally, the LBCs are derived from ERA-Interim for the evaluation with SV and DV (1989–2008, i.e., EvalSV and EvalDV) and from four GCMs for the historical (1981–2000) and future (2081–2100) periods with SV and DV, i.e., HistSV and HistDV and FutSV and FutDV, respectively, under the RCP8.5 scenario. The GCMs used in this study include the Community Earth System Model (CESM) (Hurrell et al. [31]), Geophysical Fluid Dynamics Laboratory Model (GFDL) (Team et al. [60]), Model for Interdisciplinary Research on the Climate–Earth System Model (MIROC) (Watanabe et al. [69]), and Max Planck Institute Earth System Model (MPI-ESM) (Giorgetta et al. [23]). These GCMs were chosen as they can capture present-day climate over West Africa (Cook and Vizy [12]; Roehrig et al. [54]) and their climate supports a reasonable present-day vegetation according to CLM.5-CN-DV (Yu et al. [75]). The model grid is configured using a 50-km horizontal grid spacing and 18 vertical layers from the surface to 50 hPa. The model parameterizations are the same as those used in previous studies over the same region (Alo and Wang [4]; Saini et al. [55]; Wang et al. [67]), as summarized in Table 1.

Table 1 Description of 18 different RegCM-CN-DV simulation setups

Periods

Evaluation (EvalSV, EvalDV)

1989–2008

Experimental

Historical (HistSV, HistDV)

1981–2000

Future (FutSV, FutDV)

2081–2100

Spatial resolution

50 × 50 km

Vegetation

DV

Dynamic vegetation

SV

Static vegetation

Boundary conditions

Evaluation

ERA-Interim (1.5 × 1.5°)

Experiments

CESM CCSM4

Community Earth System Model (Hurrell et al. 2013, 1.2 × 0.9°)

GFDL-ESM 2M

Geophysical Fluid Dynamics Laboratory (Team et al. 2004, 2.5 × 2.0°)

MIROC5

Model for Interdisciplinary Research on Climate Earth System Model (Watanabe et al. 2011, 1.4 × 1.4°)

MPI-ESM-MR

Max Planck Institute Earth System Model (Giorgetta et al. 2013, 1.8 × 1.8°)

Model parameterization schemes

Boundary layer

Holtslag PBL (Holtslag et al. 1990)

Cumulus convection

Emanuel scheme (Emanuel 1991)

Precipitation and cloud

Sub-grid Explicit Moisture Scheme (Pal et al. 2000)

Radiation

Community climate model 3 (Kiehl et al. 1996)

Dynamics

Mesoscale model 5 (Grell et al. 1994)

Ocean flux

Zeng scheme (Zeng et al. 1998)

Anthropogenic aerosols/interactive aerosols

Tracer model (Solmon et al. 2006; Zakey et al. 2006, 2008)

Land surface

Community Land Model 4.5-CN-DV (Oleson et al. 2010, 2013; Wang et al. 2017)

Wang et al. ([67]) extensively evaluated the RegCM-CLM-CN-DV model for simulating regional climate and ecosystems in Africa. The evaluation was performed using LBCs from the ERA-Interim (1989–2008), and with and without vegetation dynamics. Yu et al. ([75]) and Erfanian et al. ([21]) also examined the impacts of vegetation dynamics on the climate and ecosystems using multiple LBCs from past and future GCM simulations. Building upon these previous studies, this study focuses on the impacts of vegetation dynamics on the regional drought characteristic (i.e., frequency, duration, and intensity) over two focal regions of the West African domain, i.e., the Sahel and the Guinea Coast (Fig. 1).

Graph: Fig. 1 Averages of precipitation (left column), air temperature (middle column), and precipitation surplus/deficit (right column) differences for summer (June, July, and August), 1981–2000, between observations (UDEL and CRU) and ensemble historical runs (HistSV and HistDV). The top and bottom black boxes show two focal regions: the Sahel and the Guinea Coast, respectively

SPEI

Various drought indices, such as the Palmer Drought Severity Index (PDSI) (Palmer [49]) and Standard Precipitation Index (SPI) (McKee et al. [43]), have been used to assess drought events. Vicente-Serrano et al. ([65]) suggested the SPEI, which uses the deficit between precipitation and potential evapotranspiration (PET) and considers the impact of temperature, a major control on potential evapotranspiration. Since the development of SPEI, various drought studies have adopted this index (e.g., Boroneant et al. [8]; Deng [16]; Li et al. [37], [38]; Lorenzo-Lacruz et al. [39]; Paulo et al. [52]; Sohn et al. [57]; Spinoni et al. [59]; Yu et al. [73]). For example, McEvoy et al. ([42]) used SPEI as a drought index to monitor conditions over Nevada and Eastern California, suggesting that SPEI was a convenient tool to describe drought in arid regions. Recently, Diasso and Abiodun ([17]) investigated the future impacts of global warming and reforestation on drought patterns simulated by RCMs over West Africa using the SPEI.

In this study, we chose SPEI, considering the advantages of the index as suggested by Naumann et al. ([45]) and Um et al. ([63]). Drought indices that account for the impact of temperature, such as PSDI or SPEI, are generally preferred especially when future climate scenarios are involved (Begueria-Portugues et al. [7]). Furthermore, the SPEI can be quantified simply by fitting the difference between precipitation and PET to the log-logistic distribution, whereas PDSI can be quantified by simulating the calibrated water balance models.

We estimated the SPEI using the approach of Beguería and Vicente-Serrano ([6]). While precipitation is the simulated output of RegCM4.3.4-CLM-CN-DV, PET can be derived from the model outputs following the approach proposed by Thornthwaite ([61]). Mavromatis ([40]) showed that the use of simple or complex methods to calculate the PET provides similar results. For a given month (j) and year (i), the monthly water surplus or deficit (Di, j) is calculated by Eq. (1) given below:

  • Di,j=PRi,jPETi,j
  • Graph

    where PR is the precipitation and PET is the potential evapotranspiration. Then, accumulated monthly water surplus or deficits at time scale k ( Xi,jk ) is calculated based on Di, j. In this study, we chose 12 months for the time scale. As suggested in Vicente-Serrano et al. ([65]), SPEIi,jk is estimated by fitting Xi,jk to the log-logistic distribution by means of the L-moments method. In this study, a drought event is defined when the SPEI value is less than − 1.

    Results and discussion

    Model performance for the present day

    This section briefly evaluates the model performance with observed climate and vegetation characteristics. The model evaluation with the ERA-Interim for 1989–2008 (EvalSV and EvalDV in Table 1) is documented in Wang et al. ([67]), whereas the evaluation with different GCMs for 1981–2000 (HistSV and HistDV in Table 1) can be found in Erfanian et al. ([21]). Thus, we briefly evaluate the ensemble means of HistSV and HistDV in comparison with the observational data. Relative to the observational data from the University of Delaware (UDEL) and Climate Research Unit, both HistSV and HistDV (Figs. S1 and 1) follow the observed spatial patterns of precipitation and temperature. Both overestimated precipitation over the Guinea Coast and underestimated it over the Sahel. For instance, monthly mean precipitation over the Guinea Coast in summer (June, July, and August, JJA) is found to be 206.8 mm/month for UDEL with 42.6% and 46.8% overestimation for HistSV and HistDV, respectively. In general, such wet/dry biases lead to cool/warm biases in air temperature via the enhancement/reduction of evaporative cooling in the Guinea Coast/Sahel. We also investigated the precipitation surplus/deficit that is used for calculating the SPEI values to analyze the drought characteristics. We found that the differences of both HistDV and HistSV from the observational precipitation surplus/deficit follow same pattern as witnessed for precipitation in corresponding cases. To sum up, the model performs slightly better with SV than with DV in the historical runs for most cases. However, note that DV could eliminate potential inconsistancy between prescribed vegetation and climate, particularly for the future projections. Refer to Figs. S2 and S3 for the model evaluations with the ERA-Interim (EvalSV and EvalDV), which are generally consistent with HistSV and HistDV.

    With the addition of vegetation dynamics, the LAI (Fig. 2) is overestimated in the Guinea Coast, whereas it is underestimated in the Sahel (HistDV-HistSV), which is consistent with the abovementioned biases in climate conditions. Mean LAI for HistSV over the Sahel and Guinea Coast are 0.63 and 2.15, whereas those for HistDV are 0.15 and 2.61, respectively. Note that the run without vegetation dynamics (HistSV) uses the Moderate Resolution Imaging Spectroradiometer (MODIS)–based monthly varying climatological LAI values, i.e., observational data. Over the Sahel, the model underestimates the woody plants and grass coverage with a significant overestimation of bare ground area, which can be attributed to biases in the vegetation dynamics of the model (i.e., CLM-CN-DV) and physical climate of RegCM, i.e., dry bias (Wang et al. [67]; Erfanian et al. [21]). The dry/wet bias in the atmospheric forcings over the Sahel/Guinea Coast contributes to the underestimated/overestimated LAI, which then leads to additional decreases/increases in precipitation for that region (Fig. 1). Note that similar dry and wet biases were also simulated with another regional earth system model coupled with the DV model (RCA-GUESS) (Wu et al. [71]).

    Graph: Fig. 2 Averages of leaf area index (LAI) from observation (MODIS), which is used for SV runs (EvalSV, HistSV, and FutSV), and LAI simulated in DV for the evaluation run and experiment ensemble runs (EavlDV, HistDV, and FutDV) in the first row and their differences in the second row. Dotted regions show the areas that pass the two-tailed significance level with α = 0.01

    Projected future changes in droughts

    This section investigates the changes in climate, vegetation, and drought characteristics between the future and historical periods in the experimental runs (HistSV vs. FutSV and HistDV vs. FutDV). First, the projected changes in the climate conditions in the future, relative to the historical periods, are examined in Fig. 3, showing similar spatial patterns of changes in SV and DV. In DV (FutDV-HistDV), average differences in precipitation for JJA are approximately − 23.6 and 4.4 mm/month over Sahel and the Guinea Coast, respectively. Furthermore, atmospheric warming caused by the increased CO2 concentration in the future scenario leads to widespread increases (positive values) in temperatures by 4 to 5 °C for both SV and DV ensembles (second column of Fig. 3), as expected.

    Graph: Fig. 3 Difference in precipitation (mm/month, left column), air temperature (°C, middle column), and precipitation surplus/deficit (mm/month, right column) between different experimental simulations (HistSV, HistDV, FutSV, and FutDV) for summer (June, July, and August). Dotted regions show the areas that pass the two-tailed confidence level with α = 0.01

    Projected LAI changes (FutDV-HistDV of Fig. 2) result from both climate change and CO2 fertilization effects. Note that LAIs in HistSV and FutSV are identical to the observational data. In the DV ensemble, slight decreases are projected in the Sahel, while noticeable increases are projected in the Guinea Coast with an average of 1.0 m2/m2. The spatial variation of the LAI changes generally follows that of the precipitation changes (FutDV-HistDV of Fig. 3), as both are interactively simulated in the model.

    In the future, the precipitation surplus/deficit shows a general decline for both SV and DV ensembles (FutSV-HistSV and FutDV-HistDV in the right column of Fig. 3). Overwhelming decreases in the precipitation surplus (i.e., increases in deficit) are found except for the southern part of the Guinea Coast, which can be attributed to future atmospheric warming (the center column of Fig. 3). In Fig. 4, we find that the changes in the precipitation surplus/deficit lead to similar changes in drought frequencies between the future and historical periods for both SV and DV ensembles, although the spatial variations in the drought frequency changes are relatively noisy. Slight increases by 0.2% and 1.9% of the frequency in the Sahel in the future period relative to the historical period are found in the SV and DV ensembles, respectively. However, an increase of 4.1% and decrease of − 4.0% in the Guinea Coast are found in the SV and DV ensembles, respectively. The impact of precipitation increase (leading to a decrease in drought frequency) dominates the impact of temperature increases in the DV ensemble (leading to an increase in drought frequency) partly because the precipitation increase is further enhanced in the DV ensemble (FutSV-HistSV vs. FutDV-HistDV in Fig. 3).

    Graph: Fig. 4 Difference in drought frequencies between different experimental simulations (HistSV, HistDV, FutSV, and FutDV). Drought event is defined as an event with an SPEI value of less than − 1, relative to the whole period

    Impact of vegetation dynamics on simulated droughts

    Including vegetation dynamics in a land–atmosphere coupled model makes the model more complex, but is closer to the real world for climate simulation and projections. Therefore, we focus on investigating the impact of vegetation dynamics on simulated droughts (i.e., the difference between DV and SV for the future or FutDV-FutSV) in this section. Investigating the difference of LAI between DV and SV for the future period (FutDV-FutSV of Fig. 2), we find that the LAI for the DV ensemble is smaller than that of SV over the Sahel and larger over the southern part of the Guinea Coast. LAI differences between SV and DV ensembles show similar patterns both in the historical and future periods. In the historical period (HistDV vs. HistSV), LAI biases are caused by the biases from both CLM-CN-DV and RegCM as the LAI of HistSV is based on the observational data, i.e., MODIS, as previously mentioned. Yet, the lower LAI in the Sahel in the future period (FutDV vs. FutSV, average difference of − 0.51 m2/m2) is not necessarily a bias in the simulations because the future LAI in SV is assumed to be identical to historical climatological LAI as is found in the historical SV ensemble. Note that such changes in the LAI in the future do not accompany changes in the dominant vegetation types in the Sahel region. In contrast, for the Guinea Coast, the higher simulated LAI (averaged difference of 1.54 m2/m2) is associated with differences in land cover (from grasses to woody plants).

    As examined in Erfanian et al. ([21]), along with lower LAI in the Sahel, when comparing DV with SV, higher albedo, lower cooling, lower evapotranspiration, and lower precipitation are simulated. Over the region below 10° N, wetter and colder climate conditions are projected with the DV ensemble compared with the SV ensemble, resulting in a greater precipitation surplus (FutDV-Fut SV in the right column of Fig. 3). Consequently, the frequencies of drought events −10% lower over the Gulf Coast and 18% higher over the Sahel, based on the ensemble averages (FutDV-FutSV). We find that such characteristics in the ensemble averages are consistently captured in each ensemble member to different extents (Fig. S4). On average, drought frequencies are higher in the Sahel, ranging from 13 to 29% among the different LBCs and lower in the Guinea Coast, ranging from − 5 to − 20% among the different LBCs.

    The regional averages of SPEI over the two different regions present the impact of vegetation dynamics on simulated drought intensity and duration (Fig. 5). Ensemble SPEI averages and the black-lined boxes for the drought events, defined with the averaged SPEI values less than − 1, show that DV runs simulate more prolonged and more severe droughts for the historical and future periods over the Sahel, which is also shown by the negative values in the difference figures (i.e., HistDV-HistSV and FutDV-FutSV, last two rows of each panel). Specifically, the drought event is not identified in FutSV with the minimum SPEI of − 0.94, whereas 4 drought events across the total period of 42 months are characterized by an SPEI as low as − 1.61 in FutDV (first panel in Fig. 5). Such findings are generally opposite for the Guinea Coast, as shown by the positive values in the difference figures, with drought events being depicted in FutSV but not in FutDV. We also show that these findings are consistent with the ensemble members with four different LBCs. For instance, in MIROC, the first drought event is characterized by a period of 10 months, mean SPEI of − 1.32, and minimum SPEI of − 1.49 in HistSV, and with a period of 14 months, mean SPEI of − 1.75, and minimum SPEI of − 1.95 in HistDV. Furthermore, prolonged droughts with lengths of 64 months are projected in FutDV only. The data summarized in Table 2 clearly suggest that the impact of DV on drought frequency (i.e., the ratio of drought period relative to the whole study period) over the Guinea Coast shows the opposite comparison to that over the Sahel, where vegetation feedback clearly increased drought frequency. For instance, the drought frequency increases from 10.48% of HistSV to 27.16% of HistDV, and from 10.68% of FutSV to 29.01% of FutDV. We note that these findings are qualitatively consistent with the results of Diasso and Abiodun ([17]), which suggest longer and intense dry periods in the Sahel.

    Graph: Fig. 5 Monthly SPEI averaged for two focal regions of the Sahel (left column) and the Guinea Coast (right column) in ensembles and the individual member of the experimental runs (HistSV, HistDV, FutSV, and FutDV) with different LBCs of CCSM, GFDL, MIROC, and MPI-ESM. Boxes indicate the drought events defined by an SPEI value of less than − 1

    Table 2 Ensemble drought frequency (%) over the Sahel and the Guinea Coast with experimental runs (HistSV, HistDV, FutSV, and FutDV). Drought events are defined where the SPEI value is less than − 1

    HistSV

    HistDV

    FutSV

    FutDV

    Min

    Mean

    Max

    Min

    Mean

    Max

    Min

    Mean

    Max

    Min

    Mean

    Max

    Sahel

    0.32

    10.48

    22.6

    13.97

    27.16

    45.74

    1.09

    10.68

    19.54

    20.20

    29.01

    44.0

    Guinea Coast

    0.54

    22.81

    47.38

    0.21

    15.96

    43.67

    2.07

    26.80

    43.67

    0.32

    11.38

    44.0

    Finally, we perform the correlation analysis between annual maximum LAI and annual minimum SPEI over the regions for both historical and future periods (Fig. 6). Drought events are reflected in the relatively lower annual minimum SPEI because leaf growth is limited during such events. Therefore, a large portion of West Africa sees a strong positive association between the maximum LAI and minimum SPEI with DV activated. Relatively strong correlations with a coefficient up to 0.91 are found along the Sahel, which may be attributed to land–atmosphere feedbacks that are relatively strong in transition zones. Over most of the Guinea Coast, the correlation is weak or even negative, as as vegitation growth is not limited by water in the region.

    Graph: Fig. 6 Spearman's rank correlation coefficient between annual minimum LAI and annual maximum SPEI from HistDV (1981–2000; left column) and FutDV (2081–2100; right column)

    Conclusions

    In this study, we employed a drought index (i.e., SPEI) to assess the effects of vegetation dynamics on projected future droughts over West Africa, and impact of vegetation feedbacks on drought characteristics was quantified based on experiments with and without vegetation dynamics. This study suggests that when vegetation dynamics are considered, drought is prolonged and enhanced over the Sahel, and this effect is similar between present and future climates. Specifically, based on the regionally averaged ensembles of SPEIs over the Shael, no drought event is identified in FutSV with the minimum SPEI of − 0.94, whereas 4 drought events across the total period of 42 months are identified by an SPEI as low as − 1.61 in FutDV. For the Guinea Coast, vegetation dynamics has an opposit effect with less drought identified in Fut DV than in FutSV. These results are consistent over four different LBCs. Furthermore, we show that simulated vegetation density (i.e., LAI) was correlated well with annual maximun SPEI with the correlation coefficient of up to 0.91, particularly over the Sahel, a sensitive transition zone where the feedback between land and atmosphere is relatively strong.

    While most future drought characterization studies have been conducted without considering the role of interactive vegetation in climate models (e.g., Cook et al. [15]; Huang et al. [29]), this study suggests the importance of biosphere–atmosphere interactions in future drought projections. Dirmeyer et al. ([18]) suggest that land–atmosphere coupling is projected to increase in the future across most of the globe and thus the land surface is likely to play a greater role in amplifying hydrologic extremes, such as severe droughts. Our results suggest that the possible increase in severe droughts caused by climate change could be further enhanced by vegetation feedback.

    PET in this study is estimated using the Thornthwaite approach, which considers air temperature as a governing feature of PET. However, there are various other methods to calculate PET. For example, the Penman–Montieth method is more physically realistic but requires a diverse input data set (i.e., humidity, radiation coefficient, and wind speed). Van der Schrier et al. ([64]) calculated the change in the global PDSI using two distinct estimates for PET (e.g., Thornthwaite and Penman-Monteith), and found that PSDI based on two PET estimates are identical in terms of trend, average values, and classifying severe wet or dry periods. Conversely, McVicar et al. ([44]) pointed out that climatic conditions other than temperature may compensate the effect of temperature rise on drought in the future climate. These suggest that our conclusions might vary with different drought indices to some extents.

    This study points out the potential of prolonged and enhanced drought events in the Sahel by considering the vegetation dynamics, which has seldom been considered in previous studies. Furthermore, many African countries are expected to experience population growth with some of the fastest relative growth in West African countries, including Niger and Chad (Ahmadalipour et al. [1]). Combined with the high likelihood of prolonged and enhanced drought and population growth, there will likely be an increase in water shortage. This will further exacerbate the future drought risks and will present climate adaptation challenges for managing water needs in the region.

    Funding

    This study was supported by the Basic Science Research Program through the National Research Foundation of Korea, which was funded by the Ministry of Science, ICT & Future Planning (2020R1A2C2007670) and the Internationalization Infra Fund of Yonsei University (2018 Fall semester).

    Supplementary Information

    Graph: (PDF 1.50 mb).

    Publisher's note

    Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

    References 1 Ahmadalipour A, Moradkhani H, Castelletti A, Magliocca N. Future drought risk in Africa: integrating vulnerability, climate change, and population growth. Sci Total Environ. 2019; 662: 672-686 2 Akinsanola A, Zhou W (2019a) Projections of West African summer monsoon rainfall extremes from two CORDEX models. Clim Dyn 52:2017–2028 3 Akinsanola A, Zhou W (2019b) Projection of West African summer monsoon rainfall in dynamically downscaled CMIP5 models. Clim Dyn 53(1–2):81–95 4 Alo CA, Wang G. Role of dynamic vegetation in regional climate predictions over western Africa. Clim Dyn. 2010; 35: 907-922 5 Anthes RA, Hsie E-Y, Kuo Y-H. Description of the Penn State/NCAR mesoscale model version 4 (MM4). 1987: Boulder; NCAR 6 Beguería S, Vicente-Serrano SM (2013) SPEI: calculation of the standardised precipitation-evapotranspiration index R package version 1:6 7 Begueria-Portugues S, Vicente-Serrano S, Angulo-Martínez M, López-Moreno J, El Kenawy A (2010) The Standardized Precipitation-Evapotranspiration Index (SPEI): a multiscalar drought index. In: 10th EMS Annual Meeting, 10th European Conference on Applications of Meteorology (ECAM) Abstracts, held Sept. 13-17, 2010 in Zürich, Switzerland. http://meetings.copernicus.org/ems2010/. id. EMS2010-562 8 Boroneant C, Ionita M, Brunet M, Rimbu N (2011) CLIVAR-SPAIN contributions: seasonal drought variability over the Iberian Peninsula and its relationship to global sea surface temperature and large scale atmospheric circulation WCRP OSC. Climate Research in Service to Society, pp 24–28 9 Caminade C, Terray L. Twentieth century Sahel rainfall variability as simulated by the ARPEGE AGCM, and future changes. Clim Dyn. 2010; 35: 75-94 Charney J, Stone PH, Quirk WJ. Drought in the Sahara: a biogeophysical feedback mechanism. Science. 1975; 187: 434-435 Cook KH. Climate science: the mysteries of Sahel droughts. Nat Geosci. 2008; 1: 647 Cook KH, Vizy EK. Coupled model simulations of the West African monsoon system: Twentieth-and twenty-first-century simulations. J Clim. 2006; 19: 3681-3703 Cook KH, Vizy EK. Effects of twenty-first-century climate change on the Amazon rain forest. J Clim. 2008; 21: 542-560 Cook KH, Vizy EK, Launer ZS, Patricola CM. Springtime intensification of the Great Plains low-level jet and Midwest precipitation in GCM simulations of the twenty-first century. J Clim. 2008; 21: 6321-6340 Cook BI, Ault TR, Smerdon JE. Unprecedented 21st century drought risk in the American Southwest and Central Plains. Sci Adv. 2015; 1: e1400082 Deng F (2011) Global CO2 flux inferred from atmospheric observations and its response to climate variabilities. Diasso U, Abiodun BJ. Future impacts of global warming and reforestation on drought patterns over West Africa. Theor Appl Climatol. 2018; 133: 647-662 Dirmeyer PA, Jin Y, Singh B, Yan X. Trends in land–atmosphere interactions from CMIP5 simulations. J Hydrometeorol. 2013; 14: 829-849 Dixon RD, Peyrille P, Guichard F. Sahelian precipitation change induced by SST increase: the contrasting roles of regional and larger-scale drivers. Geophys Res Lett. 2019; 46; 20: 11378-11387 Emanuel KA. A scheme for representing cumulus convection in large-scale models. J Atmos Sci. 1991; 48: 2313-2329 Erfanian A, Wang G, Yu M, Anyah R. Multimodel ensemble simulations of present and future climates over West Africa: Impacts of vegetation dynamics. Journal of Advances in Modeling Earth Systems. 2016; 8: 1411-1431 Garnaud C, Sushama L, Verseghy D. Impact of interactive vegetation phenology on the Canadian RCM simulated climate over North America. Clim Dyn. 2015; 45: 1471-1492 Giorgetta MA. Climate and carbon cycle changes from 1850 to 2100 in MPI-ESM simulations for the Coupled Model Intercomparison Project phase 5. Journal of Advances in Modeling Earth Systems. 2013; 5: 572-597 Giorgi F. RegCM4: model description and preliminary tests over multiple CORDEX domains. Clim Res. 2012; 52: 7-29 Grell GA. Prognostic evaluation of assumptions used by cumulus parameterizations. Mon Weather Rev. 1993; 121: 764-787 Grell GA, Dudhia J, Stauffer DR (1994) A description of the fifth-generation Penn State/NCAR mesoscale model (MM5) Hoerling M, Hurrell J, Eischeid J, Phillips A. Detection and attribution of twentieth-century northern and southern African rainfall change. J Clim. 2006; 19: 3989-4008 Holtslag A, De Bruijn E, Pan H. A high resolution air mass transformation model for short-range weather forecasting. Mon Weather Rev. 1990; 118: 1561-1575 Huang J. Analysis of future drought characteristics in China using the regional climate model CCLM. Clim Dyn. 2018; 50: 507-525 Hulme M, Doherty R, Ngara T, New M, Lister D. African climate change: 1900-2100. Clim Res. 2001; 17: 145-168 Hurrell JW et al (2013) The community earth system model: a framework for collaborative research. Bull Amer Meteor Soc 94:1339–1360. https://doi.org/10.1175/bams-d-12-00121.1 Kamga AF, Jenkins GS, Gaye AT, Garba A, Sarr A, Adedoyin A (2005) Evaluating the National Center for Atmospheric Research climate system model over West Africa: Present-day and the 21st century A1 scenario. Journal of Geophysical Research: Atmospheres 110 Kiehl T, Hack J, Bonan B, Boville A, Briegleb P, Williamson L, Rasch J (1996) Description of the NCAR community climate model (CCM3) Kim Y, Wang G. Impact of vegetation feedback on the response of precipitation to antecedent soil moisture anomalies over North America. J Hydrometeorol. 2007; 8: 534-550 Kumar SV, Peters-Lidard CD, Eastman JL, Tao W-K. An integrated high-resolution hydrometeorological modeling testbed using LIS and WRF. Environ Model Softw. 2008; 23: 169-181 Lawrence DM et al (2011) Parameterization improvements and functional and structural advances in version 4 of the Community Land Model. J Adv Model Earth Syst 3 Li W-G, Yi X, Hou M-T, Chen H-L, Chen Z-L. Standardized precipitation evapotranspiration index shows drought trends in China. Chin J Eco-Agric. 2012; 20: 643-649 Li W, Hou M, Chen H, Chen X. Study on drought trend in south China based on standardized precipitation evapotranspiration index. Journal of Natural Disasters. 2012; 21: 84-90 Lorenzo-Lacruz J, Vicente-Serrano SM, López-Moreno JI, Beguería S, García-Ruiz JM, Cuadrat JM. The impact of droughts and water management on various hydrological systems in the headwaters of the Tagus River (central Spain). J Hydrol. 2010; 386: 13-26 Mavromatis T. Drought index evaluation for assessing future wheat production in Greece. Int J Climat. 2007; 27: 911-924 Maynard K, Royer J-F, Chauvin F. Impact of greenhouse warming on the West African summer monsoon. Clim Dyn. 2002; 19: 499-514 McEvoy DJ, Huntington JL, Abatzoglou JT, Edwards LM. An evaluation of multiscalar drought indices in Nevada and eastern California. Earth Interact. 2012; 16: 1-18 McKee TB, Doesken NJ Kleist J The relationship of drought frequency and duration to time scales. In: Proceedings of the 8th Conference on Applied Climatology, 1993. vol 22. American Meteorological Society, Boston, MA, pp 179–183 McVicar TR. Global review and synthesis of trends in observed terrestrial near-surface wind speeds: implications for evaporation. J Hydrol. 2012; 416: 182-205 Naumann G, Dutra E, Barbosa P, Pappenberger F, Wetterhall F, Vogt J. Comparison of drought indicators derived from multiple data sets over Africa. Hydrol Earth Syst Sci. 2014; 18: 1625 Oleson KW et al. (2010) Technical description of version 4.0 of the Community Land Model (CLM) Oleson K, Lawrence D et al (2013) Technical description of version 4.5 of the community land model (CLM) NCAR Earth Systems Laboratory, Climate and Global Dynamics Division ISSN Print Edition:2153–2397 Pal JS, Small EE, Eltahir EA. Simulation of regional-scale water and energy budgets: representation of subgrid cloud and precipitation processes within RegCM. J Geophys Res Atmos. 2000; 105: 29579-29594 Palmer WC (1965) Meteorological drought. Research Paper No. 45. Washington, DC: US Department of Commerce Weather Bureau:59 Patricola C, Cook KH (2008) Atmosphere/vegetation feedbacks: A mechanism for abrupt climate change over northern Africa. J Geophys Res Atmos 113 Patricola CM, Cook KH. Northern African climate at the end of the twenty-first century: an integrated application of regional and global climate models. Clim Dyn. 2010; 35: 193-212 Paulo A, Rosa R, Pereira L. Climate trends and behaviour of drought indices based on precipitation and evapotranspiration in Portugal. Nat Hazards Earth Syst Sci. 2012; 12: 1481-1491 Peyrillé P, Lafore JP, Boone A. The annual cycle of the West African monsoon in a two-dimensional model: mechanisms of the rain-band migration. Q J R Meteorol Soc. 2016; 142; 696: 1473-1489 Roehrig R, Bouniol D, Guichard F, Hourdin F, Redelsperger J-L. The present and future of the west African monsoon: a process-oriented assessment of CMIP5 simulations along the AMMA transect. J Clim. 2013; 26: 6471-6505 Saini R, Wang G, Yu M, Kim J. Comparison of RCM and GCM projections of boreal summer precipitation over Africa. J Geophys Res Atmos. 2015; 120: 3679-3699 Smith B, Samuelsson P, Wramneby A, Rummukainen M. A model of the coupled dynamics of climate, vegetation and terrestrial ecosystem biogeochemistry for regional applications. Tellus Ser A Dyn Meteorol Oceanogr. 2011; 63: 87-106 Sohn SJ, Ahn JB, Tam CY. Six month–lead downscaling prediction of winter to spring drought in South Korea based on a multimodel ensemble. Geophys Res Lett. 2013; 40: 579-583 Solmon F, Giorgi F, Liousse C. Aerosol modelling for regional climate studies: application to anthropogenic particles and evaluation over a European/African domain. Tellus Ser B Chem Phys Meteorol. 2006; 58: 51-72 Spinoni J. An overview of drought events in the Carpathian region in 1961–2010. Adv Sci Res. 2013; 10: 21-32 Team GGAMD. The new GFDL global atmosphere and land model AM2–LM2: evaluation with prescribed SST simulations. J Clim. 2004; 17: 4641-4673 Thornthwaite CW. An approach toward a rational classification of climate. Geogr Rev. 1948; 38: 55-94 Tiedtke M. A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon Weather Rev. 1989; 117: 1779-1800 Um MJ, Kim Y, Kim J. Evaluating historical drought characteristics simulated in CORDEX East Asia against observations. Int J Climatol. 2017; 37: 4643-4655 Van der Schrier G, Jones P, Briffa K (2011) The sensitivity of the PDSI to the Thornthwaite and Penman-Monteith parameterizations for potential evapotranspiration. J Geophys Res Atmos 116 Vicente-Serrano SM, Beguería S, López-Moreno JI. A multiscalar drought index sensitive to global warming: the standardized precipitation evapotranspiration index. J Clim. 2010; 23: 1696-1718 Wang G, Eltahir EA. Ecosystem dynamics and the Sahel drought. Geophys Res Lett. 2000; 27: 795-798 Wang G, Yu M, Pal JS, Mei R, Bonan GB, Levis S, Thornton PE. On the development of a coupled regional climate–vegetation model RCM–CLM–CN–DV and its validation in tropical Africa. Clim Dyn. 2016; 46: 515-539 Wang D, Wang G, Parr DT, Liao W, Xia Y, Fu C. Incorporating remote sensing-based ET estimates into the community land model version 4.5. Hydrol Earth Syst Sci. 2017; 21: 3557-3577 Watanabe S. MIROC-ESM: model description and basic results of CMIP5-20c3m experiments. Geosci Model Dev Discuss. 2011; 4: 1063-1128 Wramneby A, Smith B, Samuelsson P (2010) Hot spots of vegetation-climate feedbacks under future greenhouse forcing in Europe. J Geophys Res Atmos 115 Wu M. Vegetation–climate feedbacks modulate rainfall patterns in Africa under future climate change. Earth Syst Dynam. 2016; 7: 627-647. 10.5194/esd-7-627-2016 Xue Y, Boone A, Taylor CM (2012) Review of recent developments and the future prospective in West African atmosphere/land interaction studies. Int J Geophys 2012 Yu M, Li Q, Hayes MJ, Svoboda MD, Heim RR. Are droughts becoming more frequent or severe in China based on the standardized precipitation evapotranspiration index: 1951–2010?. Int J Climatol. 2014; 34: 545-558 Yu M, Wang G, Parr D, Ahmed KF. Future changes of the terrestrial ecosystem based on a dynamic vegetation model driven with RCP8. 5 climate projections from 19 GCMs. Clim Chang. 2014; 127: 257-271 Yu M, Wang G, Pal JS. Effects of vegetation feedback on future climate change over West Africa. Clim Dyn. 2016; 46: 3669-3688 Zakey A, Solmon F, Giorgi F. Implementation and testing of a desert dust module in a regional climate model. Atmos Chem Phys. 2006; 6: 4687-4704 Zakey A, Giorgi F, Bi X (2008) Modeling of sea salt in a regional climate model: Fluxes and radiative forcing. J Geophys Res Atmos 113 Zeng X, Zhao M, Dickinson RE. Intercomparison of bulk aerodynamic algorithms for the computation of sea surface fluxes using TOGA COARE and TAO data. J Clim. 1998; 11: 2628-2644 Zhang W, Jansson C, Miller PA, Smith B, Samuelsson P. Biogeophysical feedbacks enhance the Arctic terrestrial carbon sink in regional Earth system dynamics. Biogeosciences. 2014; 11: 5503-5519

    By Muhammad Shafqat Mehboob; Yeonjoo Kim; Jaehyeong Lee; Myoung-Jin Um; Amir Erfanian and Guiling Wang

    Reported by Author; Author; Author; Author; Author; Author

    Titel:
    Projection of vegetation impacts on future droughts over West Africa using a coupled RegCM-CLM-CN-DV
    Autor/in / Beteiligte Person: Um, Myoung-Jin ; Erfanian, Amir ; Lee, Jaehyeong ; Wang, Guiling ; Muhammad Shafqat Mehboob ; Kim, Yeonjoo
    Link:
    Zeitschrift: Climatic Change, Jg. 163 (2020-11-01), S. 653-668
    Veröffentlichung: Springer Science and Business Media LLC, 2020
    Medientyp: unknown
    ISSN: 1573-1480 (print) ; 0165-0009 (print)
    DOI: 10.1007/s10584-020-02879-z
    Schlagwort:
    • Atmospheric Science
    • Global and Planetary Change
    • 010504 meteorology & atmospheric sciences
    • 0208 environmental biotechnology
    • Biosphere
    • 02 engineering and technology
    • Vegetation
    • 01 natural sciences
    • 020801 environmental engineering
    • Atmosphere
    • Evapotranspiration
    • Climatology
    • Environmental science
    • Climate model
    • Precipitation
    • Leaf area index
    • Projection (set theory)
    • 0105 earth and related environmental sciences
    Sonstiges:
    • Nachgewiesen in: OpenAIRE
    • Rights: CLOSED

    Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

    oder
    oder

    Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

    oder
    oder

    Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

    xs 0 - 576
    sm 576 - 768
    md 768 - 992
    lg 992 - 1200
    xl 1200 - 1366
    xxl 1366 -