ArticlePDF Available

Climate and Vegetation in the Middle East: Interannual Variability and Drought Feedbacks

Authors:

Abstract and Figures

The Euphrates Plain (EP) experiences large interannual variability in vegetation cover, especially in areas of marginal rain-fed agriculture. Vegetation in this region is primarily limited by available soil moisture, as determined by winter precipitation, spring precipitation, and air temperature. Satellite analyses indicate that the springtime normalized difference vegetation index (NDVI) is negatively correlated with surface albedo, and that interannual variability in albedo in the EP produces an estimated forcing on the radiation balance that peaks at 16.0 W m−2 in May. Simulations with a regional climate model indicate that surface energy fluxes during a drought year (1999) differed substantially from those during a year with normal precipitation (2003). These differences were geographically specific, with the EP exhibiting increased albedo and decreased sensible heat flux while the neighboring Zagros Plateau region showed no albedo effect, a large increase in sensible heat flux, and an offsetting reduction in latent heat flux. In both the EP and the Zagros there was a potential for positive feedbacks on temperature and drought in late spring, though the most likely feedback mechanisms differed between the two regions: in the EP surface brightening leads to cooling and reduced turbulent heat flux, while in the Zagros region reduced latent heat flux leads to warming and a deepening of the planetary boundary layer.
Content may be subject to copyright.
Climate and Vegetation in the Middle East: Interannual Variability and
Drought Feedbacks
BENJAMIN F. ZAITCHIK,* JASON P. EVANS,ROLAND A. GEERKEN, AND RONALD B. SMITH
Department of Geology and Geophysics, Yale University, New Haven, Connecticut
(Manuscript received 23 May 2006, in final form 9 November 2006)
ABSTRACT
The Euphrates Plain (EP) experiences large interannual variability in vegetation cover, especially in areas
of marginal rain-fed agriculture. Vegetation in this region is primarily limited by available soil moisture, as
determined by winter precipitation, spring precipitation, and air temperature. Satellite analyses indicate that
the springtime normalized difference vegetation index (NDVI) is negatively correlated with surface albedo,
and that interannual variability in albedo in the EP produces an estimated forcing on the radiation balance
that peaks at 16.0 W m
2
in May.
Simulations with a regional climate model indicate that surface energy fluxes during a drought year (1999)
differed substantially from those during a year with normal precipitation (2003). These differences were
geographically specific, with the EP exhibiting increased albedo and decreased sensible heat flux while the
neighboring Zagros Plateau region showed no albedo effect, a large increase in sensible heat flux, and an
offsetting reduction in latent heat flux. In both the EP and the Zagros there was a potential for positive
feedbacks on temperature and drought in late spring, though the most likely feedback mechanisms differed
between the two regions: in the EP surface brightening leads to cooling and reduced turbulent heat flux,
while in the Zagros region reduced latent heat flux leads to warming and a deepening of the planetary
boundary layer.
1. Introduction
Semiarid regions are subject to regular seasonal dry-
ness and large interannual variability in precipitation.
This results in variable vegetation cover on annual and
interannual time scales, as both natural ecosystems and
nonirrigated crops rely on soil moisture derived from
seasonal rains or springtime snowmelt (Baldocchi et al.
2004; Dall’Olmo and Karnieli 2002; Evans and Geerken
2004; Weiss et al. 2004). Opportunistic annual species
green up rapidly in response to wetting of the soil sur-
face, and their vigor is primarily related to recent rain-
fall events. Winter crops and perennial vegetation have
access to deeper reserves of soil moisture. Growth of
these vegetation types depends on the precipitation
pattern over weeks and months, on evaporative de-
mand, and, for some regions, on temperature con-
straints.
Climate-induced variability in semiarid vegetation is
a matter of both ecological interest and economic con-
cern, as strong sensitivity to climate can result in rapid
land use change (Vanacker et al. 2005) and vulnerabil-
ity to human-induced degradation (Evans and Geerken
2004). Over longer time scales, relatively small shifts in
background climate may have a substantial influence
on the distribution of ecosystems and, perhaps, the vi-
ability of agricultural and pastoral systems (deMenocal
2001; Hole 1994; Weiss and Bradley 2001). Interest in
the climate sensitivities of semiarid vegetation—par-
ticularly crops and rangelands—is evident in the large
body of research devoted to characterizing the relation-
ships between precipitation, soil type, land manage-
ment, and vegetation growth in water-stressed regions
(e.g., Archer et al. 1995; Kremer et al. 1996; Lane et al.
1998; LeHouerou 1996).
There is also a substantial literature concerned with
vegetation feedbacks on climate in semiarid regions. In
an effort to understand the role of land–atmosphere
interactions during the extended Sahel drought of the
early 1970s, Charney (1975) hypothesized a negative
albedo–precipitation feedback in which dry conditions
* Current affiliation: Hydrologic Sciences Branch, NASA God-
dard Space Flight Center, Greenbelt, Maryland.
Corresponding author address: Benjamin F. Zaitchik, Hydro-
logic Sciences Branch, NASA Goddard Space Flight Center,
Code 614.3, Greenbelt, MD 20771.
E-mail: bzaitchik@hsb.gsfc.nasa.gov
3924 JOURNAL OF CLIMATE VOLUME 20
DOI: 10.1175/JCLI4223.1
© 2007 American Meteorological Society
JCLI4223
lead to surface brightening, a decrease in available en-
ergy, and reduced convection. This, in turn, leads to
vegetation die-back and increased albedo. Charneys
hypothesis has informed numerous studies on albedo
precipitation feedbacks, including detailed field cam-
paigns (Eltahir 1998; Small and Kurc 2003), satellite
analyses (Brunsell 2006; Courel et al. 1984), and global-
scale modeling exercises (Laval and Picon 1986; Sud
and Fennessy 1982). This line of investigation has re-
vealed important interactions between vegetation, al-
bedo, and climate. The nature of these interactions de-
pends on scale and regional context, and the forcing
mechanisms are not always obvious.
In addition to albedo effects, vegetation is thought to
influence the atmosphere through a number of struc-
tural and physiological mechanisms. Vegetation status
has a significant impact on surface roughness, and the
decrease in roughness associated with drought, land
clearing, or overgrazing can lead to decreases in aero-
dynamic conductivity to surface heat fluxes, limiting
energy transport to the atmosphere, and reducing con-
vection (Sud and Smith 1985; Zheng et al. 2002). Veg-
etation also exerts direct control on latent heat flux
from the surface. An active vegetation cover tends to
increase latent heat flux, due to increased soil infiltra-
tion and subsequent transpiration of otherwise unavail-
able moisture. Increased latent heat flux both humidi-
fies the planetary boundary layer (PBL) and increases
moist static energy (MSE) of near-surface air, increas-
ing the potential for precipitation (Eltahir 1998; Shukla
and Mintz 1982; Sud and Fennessy 1984). Such coupling
between surface conditions and atmospheric processes
is thought to be particularly strong in semiarid regions
(Koster et al. 2004).
This study takes advantage of recent interannual cli-
matic variability to characterize the climatic sensitivi-
ties of vegetation in the Middle East and to investigate
the potential for landatmosphere feedbacks in the re-
gion. In the first half of this study (sections 24) time
series data from the Advanced Very High Resolution
Radiometer (AVHRR), Systeme pour lObservation
de la Terre (SPOT)-Vegetation, and Moderate Imaging
Spectrometer (MODIS) satellite sensors are used to
describe seasonal and interannual patterns in vegeta-
tion and albedo. Data sources and processing are de-
tailed in section 2. In section 3 the phenology of major
land cover types is described, interannual variability
over the period 19812001 is quantified using AVHRR,
and the primary climatic drivers of vegetation variabil-
ity are identified. In section 4 satellite data are used to
test the hypothesis that springtime vegetation growth
has an impact on surface albedo and the surface energy
balance in both spring and summer.
The second half of the study focuses on the years
1999 and 2003a drought year and a nondrought year,
respectivelyfor detailed analysis of landatmosphere
processes during drought, including a functional com-
parison of the lowlands of the Euphrates Plain (EP)
and the uplands of the neighboring Zagros Plateau.
This is accomplished by applying the PSU/NCAR MM5
regional climate model (Dudhia 1993; Grell et al. 1994)
in full-year simulations for both 1999 and 2003, with
satellite-derived datasets used to provide lower bound-
ary information on vegetation cover and surface al-
bedo. Section 5 provides details on the MM5 simula-
tions. Results of the MM5 experiments are presented in
section 6, and in section 7 the potential for feedbacks on
cloud cover and precipitation are discussed. General
conclusions are offered in section 8.
2. Data processing
a. Satellite data
The AVHRR satellite, with its 20-yr data record
(19812001) and reasonably high spatial resolution (8
km), provides an excellent tool for the analysis of re-
gional vegetation. AVHRR 10-day composites of sur-
face reflectance and maximum normalized differential
vegetation index (NDVI) were downloaded from the
NASA Distributed Active Archive Center (DAAC).
Next, following a method described by Los (1993), time
series AVHRR data were calibrated against three fairly
time-invariant desert targets located in the Saudi Ara-
bian desert. The method removes effects of sensor deg-
radation and corrects for calibration differences be-
tween different sensor systems. Where indicated, these
corrected 10-day images were averaged to provide
mean monthly values.
Broadband shortwave albedo (
AVHRR
) was calcu-
lated from AVHRR reflectance data (r
1
, r
2
) using the
empirical relationship of Liang et al. (2002):
AVHRR
⫽⫺0.3376r
1
2
0.2707r
2
2
0.7074r
1
r
2
0.2915r
1
0.5256r
2
0.0035,
where r
1
is red reflectance and r
2
is reflectance in the
near-infrared. This nonlinear formula performed quite
well in initial validation (Liang et al. 2002). In the
present study, reliability of the estimate was confirmed
by comparison with the sophisticated albedo estimates
provided by MODIS on board the Terra satellite. Dur-
ing the 19 months of sensor overlap, AVHRR yielded
albedo estimates for the study area that had similar
variability to those of the MODIS Bidirectional Reflec-
tance Distribution Function (BRDF) albedo product
(mod43B3) (correlation for average EP albedo r
1AUGUST 2007 Z A I TCHIK ET AL. 3925
0.88). AVHRR-derived albedo was slightly, but consis-
tently, lower than MODIS values by an average of 0.07
for the domain of interest. This offset would not be
expected to affect the statistical analyses applied in this
study.
For MM5 experiments, both albedo and green veg-
etation fraction were derived from 10-day composite
images from the SPOT-Vegetation sensor. SPOT-
Vegetation provides high quality, corrected surface re-
flectance data (Maisongrande et al. 2004), and the sen-
sor was active in both 1999 and 2003, the two years of
analysis in this study. Green vegetation fraction (f
g
)
was calculated from SPOT NDVI using the linear veg-
etation index to percent cover conversion of Gutman
and Ignatov (1998):
f
g
NDVI NDVI
0
NDVI
NDVI
0
.
Minimum and maximum NDVI (NDVI
0
and NDVI
)
values were set to the minimum and maximum 10-day
values for the study region over the available time se-
ries of SPOT-Vegetation data (19992005). Broadband
albedo (
SPOT
) was calculated using the Liang et al.
(2002) empirical reflectance-to-albedo relationship:
SPOT
0.3512r
1
0.1629r
2
0.3415r
3
0.1651r
4
,
where r
14
are reflectance in the blue, green, red, and
near-infrared bands. SPOT-derived estimates of albedo
were confirmed to be statistically similar to the MODIS
BRDF albedo product over areas of comparison.
Information on land cover type is used only qualita-
tively in this study. Where possible, land use identifica-
tion was accomplished by field visit. For other areas
land cover type was determined using a satellite-de-
rived Fourier filtered cycle similarity (FFCS) technique
detailed in previous papers (Evans and Geerken 2006;
Geerken et al. 2005b).
b. Climate data
Observational records of precipitation and tempera-
ture were extracted from the global summary of month
observations collected at the Climate Prediction Center
(CPC) of the National Centers for Environmental Pre-
diction (NCEP). The study includes 824 stations, at
least 600 of which reported in each month. Station data
were interpolated using a Cressman analysis approach
with variable radius of influence. A 5°Ckm
1
lapse rate
correction was applied for interpolated temperatures,
but no topographic correction was made to precipita-
tion.
3. Vegetation types and variability in the Middle
East
The Middle East is a predominantly semiarid region
that contains a strong north to south precipitation gra-
dient (Fig. 1). Humid regions of Turkey and Transcau-
casia receive more than 1000 mm precipitation per
year, while the deserts south of the Euphrates River
receive 100 mm yr
1
or less. Interannual variability ex-
ceeds mean annual precipitation throughout the south-
FIG. 1. The Middle East: (a) Major geographic features and topography (m); (b) mean annual precipitation from FAO reporting
weather stations (19401973) interpolated using Cressman distance weighting (mm).
3926 JOURNAL OF CLIMATE VOLUME 20
ern portion of the region. This variability is of particu-
lar interest because it coincides with the southern limit
of the historical Fertile Crescent agricultural zone and
important rangelands of the Euphrates Plain. Present-
day variability has a significant impact on crop yields
and range productivity (Schmidt and Karnieli 2000;
Weiss et al. 2001), and variability on longer time scales
appears to be associated with the rise and fall of early
civilizations (Weiss and Bradley 2001).
The north to south precipitation gradient is accom-
panied by an ecological gradient ranging from temper-
ate forests and warm season agriculture in the north to
winter crops, grasslands, and eventually shrublands and
desert in the south. These land cover types differ in
vegetation density and phenology (Fig. 2a). Forests in
the highlands of Turkey are characterized by seasonally
high NDVI, with vegetation intensity peaking in early
summer and fading gradually in late summer and au-
tumn (Fig. 2b). Rain-fed agriculture in the northern
portion of the domain follows a similar pattern, return-
ing negative NDVI values in the winter (indicative of
snow cover), a rapid green-up in late spring, and rela-
tively high NDVI throughout the long summer growing
season. Agricultural phenologies in the Fertile Crescent
are quite different. Here agriculture is limited by sum-
mertime dryness rather than wintertime frost, and the
NDVI of rain-fed crops peaks in late spring (Fig. 2b).
Harvest of field crops in this region takes place in May
and June. Orchard crops such as olive also have their
phenologic peak in springtime, but the active growth
period is both longer and more moderate (Fig. 2b).
Along the river valleys and in regions of intense canal
or groundwater irrigation a third field crop phenology
is found, indicative of double-cropping practices (Fig.
2c). In these situations the winter crop may receive
supplementary irrigation and the summer crop is en-
tirely dependent on irrigation water. It should be noted
that in AVHRR-based analysis many agricultural fields
are contained within a single pixel, so a double crop
phenology actually represents a composite of winter
cropping, summer cropping, and some fields with active
double cropping. Outside of irrigated areas, land cover
south of the Fertile Crescent, east of the Caspian Sea,
and in much of Iran is limited to sparse grasslands and
shrubs. These lands are utilized seasonally for grazing,
but they have extremely limited vegetation cover for
much of the year. Phenologically, grasses and shrubs
both green up in response to winter rains, peaking in
March or April (Fig. 2c). Senescence comes rapidly in
late spring, though this pattern varies with grazing pres-
sure and the density of shrub cover, as shrubs have
access to deeper moisture reserves and can stay green
FIG. 2. Vegetation intensity in the Middle East. (a) Map of mean annual NDVI maximum, based on monthly averaged AVHRR data,
19812001. Color hue indicates the month of peak NDVI and color saturation indicates the magnitude of the NDVI peak. (b), (c)
Phenology of representative pixels for major land cover types, based on SPOT 10-day NDVI composites for 2003.
1A
UGUST 2007 Z A I TCHIK ET AL. 3927
Fig 2 live 4/C
well into the dry season (Geerken et al. 2005b). Finally,
much of the region is essentially barren, with no distinct
NDVI phenology (Fig. 2c).
For the TigrisEuphrates watershed, interannual
variability in vegetation is greatest along the southern
margin of the Fertile Crescent, an area which includes
both rain-fed agriculture and semiarid rangelands (Fig.
3). In wet years this area is lush with grain crops and
forage material, but in a drought year both crops
and range vegetation can fail entirely. In addition to
this swath of climate-driven variability, Fig. 3a shows
hot spots of anthropogenic effects. These include
the marshes of Mesopotamia, which experienced sig-
nificant human modification over the study period
(Nielsen and Adriansen 2005).
The nature of climate-driven variability can be ex-
plored by analyzing AVHRR data in conjunction with
meteorological data. For each pixel in the study region,
independent linear correlations were calculated be-
tween the annual maximum NDVI in AVHRR data
(n 20 yr) and six variables that describe weather
conditions in each hydrologic year: total precipitation
in early winter (NovemberDecember), late winter
(JanuaryFebruary), and spring (MarchApril), as
well as average air temperature for the same three
periods (data are described in section 2). Figure 4
maps the results of this analysis for all pixels with rea-
sonably substantial vegetation coverage (mean annual
NDVI
max
0.12) and significant climate sensitivity
(linear correlation between NDVI
max
and at least one
of the six climate variables significant at
0.1). Col-
ors on the map indicate the climate variable for which
the coefficient of linear correlation was largest for a
given pixel. In highland regions of northern Iran, for
example, we find that variability in NDVI correlates
most strongly with wintertime temperatures (Fig. 4,
area A): low wintertime temperatures in this area are
associated with strong vegetation growth in spring. The
negative correlation between wintertime temperature
and NDVI
max
can be attributed to the importance of
subzero temperatures for the development of a winter
snowpack, which is the primary source of springtime
soil moisture.
Moving downslope and southward along the precipi-
tation gradient, vegetation growth in the Fertile Cres-
cent and rangelands of the EP correlates more strongly
with precipitation than with temperature (Fig. 4, area
B). The mediating factor is, again, soil moisture, but
in this area snow is not a significant factor in the
water balance of nonirrigated lands. Instead, soil mois-
ture is replenished by winter and spring rains. In
areas where soil moisture storage is sufficient, deep-
rooted woody plants or winter crops are able to ac-
cess moisture that infiltrated during winter rain events
well into the spring, leading to a strong correlation be-
tween JanuaryFebruary precipitation and NDVI
max
.
In areas with shallow soils, or those dominated by op-
portunistic annual grass ecosystems, NDVI
max
corre-
lates most strongly with MarchApril precipitation, the
period coinciding with the annual vegetation maximum.
FIG. 3. NDVI variability, as captured by AVHRR monthly composites. (a) Map of the standard deviation in annual
maximum NDVI value, 19822001. Box indicates the location of the Aleppo Steppe field experiment. (b) Subset of the
AVHRR NDVI time series for a pixel within the Aleppo Steppe, giving an impression of typical annual and interannual
variability at a rangeland location.
3928 JOURNAL OF CLIMATE VOLUME 20
Sensitivity to precipitation is also dominant in the
southern Zagros Plateau (area C), where vegetation
includes hillslope grasses and several large agricultural
areas. Somewhat surprisingly, a portion of arid north-
central Saudi Arabia also exhibits sensitivity to precipi-
tation (area D). The most striking vegetation features
in this area are large cropped fields that are supported
almost exclusively by groundwater irrigation (FAO
1997). The area also includes several large wadis, how-
ever, and it is possible that the detected climate sensi-
tivity results from the response of natural vegetation to
rare wadi floods.
Finally, some areas show a significant positive corre-
lation between air temperature and NDVI
max
. These
areas fall in two regions. First, along the Mediterranean
coast and in south-central Turkey (area E) the correla-
tion with temperature reflects the fact that cold winters
limit the growth of winter crops and orchards, which
normally peak early in the year (Fig. 2a). The area is
humid relative to the interior of the Middle East (mean
annual precipitation of 350700 mm yr
1
), so the cor-
relation between vegetation and precipitation is less im-
portant. The second region of positive temperature sen-
sitivity is Mesopotamia (area F). Land use in this region
includes vast areas of irrigated agriculture, so vegeta-
tion growth is decoupled from local precipitation. In-
stead, warm winters allow for healthy growth of the
cold season crop and a higher peak NDVI.
4. Albedo and the surface energy balance
Vegetation is known to impact surface albedo in sev-
eral ways. The most direct effect is that photosyntheti-
cally active vegetation is dark in the visible range of the
electromagnetic spectrum, where incoming solar radia-
tion is greatest. This means that live plant material re-
flects less solar radiation than most semiarid soils
(Charney et al. 1977). Senescent or dormant vegetation
also influences surface albedo. Dry herbaceous mate-
rial is relatively bright, so the presence of leaf litter over
a wet or dark soil can cause an increase in albedo.
FIG. 4. Linear correlation between annual maximum NDVI and variables of winter/spring climate.
Climate parameters are average air temperature and accumulated precipitation for NovDec, JanFeb,
and MarApr in the hydrologic year of the NDVI maximum. Colors indicate the climate variable of
greatest correlation for all pixels with mean annual NDVI
max
0.12 and linear correlation is significant
at
0.1. Mean correlation coefficient for each class is indicated on the label bar. Dashed lines indicate
regions of coherent correlation type discussed in the text. Climate data are interpolated from the CPC
weather station network; NDVI is derived from AVHRR, 19812001.
1A
UGUST 2007 Z A I TCHIK ET AL. 3929
Fig 4 live 4/C
Woody material, on the other hand, tends to be darker
than dry leaves. Dormant trees and shrubs also cast
shadow, reducing the total solar radiation incident on
the soil surface. If the background soil is brighter than
the woody material itself, then this effect will lead to a
decrease in effective albedo and an increase in ab-
sorbed radiation.
Over 20 years of AVHRR data, there was a consis-
tent negative correlation between spring NDVI in the
EP and albedo in both spring and summer; that is,
strong springtime vegetation made for a darker surface
throughout the summer months. This correlation is
strongest in the rangelands and is observed as a non-
significant tendency for the EP on the whole (EP av-
erage statistics: r ⫽⫺0.44, p 0.13: Fig. 5a). There are
three reasons for this correlation in the rangelands.
First, at the resolution of an AVHRR pixel, the NDVI
signal can be dominated by wadis and depressions that
remain green throughout the summer in years with
plentiful runoff. These concentrated areas of dark veg-
etation reduce the reflectance values recorded for the
entire pixel. Second, certain succulent shrubs that are
unpalatable to livestock can retain green vegetation
well into the dry season (Geerken and Ilaiwi 2004), and
these shrubs are most green in years with good precipi-
tation. Third, summer albedo is negatively correlated
with summer NDVI due to the structural effects of
woody vegetation described above. Thus, the albedo
effect can persist even after all green vegetation has
senesced. This third phenomenon was confirmed in a
field experiment in the Aleppo Steppe: regular mea-
surements with a portable spectrometer indicated that
high spring NDVI was associated with low summer al-
bedo across a variety of rangeland environments, even
after differences in NDVI had dropped to near zero
(Figs. 5b and 5c; Geerken et al. 2005a).
To carry the analysis a step further, it is possible to
calculate the influence of interannual variability in sur-
face albedo on the absorption of shortwave radiation,
using AVHRR estimates of albedo and the 40-yr NCEP
NCAR Reanalysis Project (NNRP: Kalnay et al. 1996)
estimates of surface-incident shortwave radiation. Ac-
cording to this calculation, the interannual standard de-
viation in surface albedo is associated with a forcing on
FIG. 5. NDVI and albedo. (a) Scatterplot of June albedo against March NDVI for 20 years of AVHRR data,
averaged for all pixels in the MEP. (b), (c) Relationship between NDVI and broadband albedo for steppe
vegetation (shrub and grass), measured using a portable spectrometer in the Aleppo Steppe (Syria: Fig. 3) in 2001.
(b) Annual maximum albedo plotted against annual maximum NDVI for three subplots at three different field sites
in the steppe. Lines connect the three subplots measured within each field site. Subplots were selected to capture
a representative range of vegetation cover. (c) Weekly NDVI and albedo derived from spectrometer readings for
the three subplots at one study site [the solid diamond () site in (b)].
3930 JOURNAL OF CLIMATE VOLUME 20
the surface radiation balance that exceeds 10.0 W m
2
in every month from March through August, averaged
monthly for the EP. This forcing peaks at 16.0 W m
2
for the month of May (Fig. 6). The effect is strongest in
areas with large interannual variability in albedo and, it
must be emphasized, is an average that includes much
higher daytime values.
To put this forcing in context, the radiative forcing
associated with a doubling of atmospheric CO
2
is on the
order of 3.7 W m
2
(Myhre et al. 1998). In a regional
climate modeling study, Schär et al. (2004) found that a
17.1 W m
2
increase in net surface radiation was suffi-
cient to produce feedbacks on precipitation in sensitive
regions. Clearly, interannual variability in albedo in the
EP has the potential to have a considerable impact on
local temperatures and, potentially, hydrometeorology
on the regional scale. These effects are investigated in
greater detail in sections 5 and 6.
5. MM5 simulations
a. 1999 and 2003
Between 1998 and 2001 much of the Middle East
experienced an extended drought (DePauw 2004). The
driest year was 1999, when spatially averaged annual
precipitation in the EP was only 91.5 mm [CPC Merged
Analysis of Precipitation (CMAP): Xie and Arkin
(1996)]. The rainy season of 2003 represented a return
to normal hydrologic conditions. The year was not uni-
formly wet throughout the Middle East, but large por-
tions of the region received precipitation that was at or
above the long-term average. The EP received 163.2
mm of precipitation, and precipitation was greater in
2003 than in 1999 for every month of the rainy season
(Fig. 7a). NDVI was correspondingly higher across the
EP and most of the Zagros Plateau.
b. MM5 experiments
The application of MM5 and the NCEPOregon
State UniversityU.S. Air ForceHydrologic Research
Laboratory (NOAH) land surface model (LSM) to this
region has been described elsewhere (Evans and Smith
2005; Zaitchik et al. 2005). The standard, limited-area,
nonhydrostatic PSUNCAR MM5 is used (Dudhia
1993; Grell et al. 1994), implemented with the Reisner
mixed-phase explicit moisture scheme (Reisner et al.
1998), the Rapid Radiative Transfer Model (RRTM)
radiation scheme (Mlawer et al. 1997), Grells cumulus
scheme (Grell et al. 1994), and the Medium-Range
Forecast (MRF) models planetary boundary layer
scheme (Hong and Pan 1996). The MRF scheme is a
first-order, nonlocal scheme optimized to represent
large-eddy turbulence in a well-mixed PBL. In a com-
parison study of MM5 PBL schemes performed for the
southwestern United Statesa region that contains
both intense deserts and significant topography, much
like the Middle Eastit was found that first-order clo-
sure schemes, including MRF, produced the highest ac-
curacy predictions of PBL depth, CAPE, and vertical
profiles of temperature and mixing ratio (Bright and
Mullen 2002).
The NOAH LSM is used in its coupled form. Noah is
a direct descendent of the Oregon State University
(OSU) LSM (Mahrt and Pan 1984; Mahrt and Ek 1984;
Pan and Mahrt 1987), a sophisticated land surface
model that has been extensively validated in both
coupled and uncoupled studies (Chen and Mitchell
1999; Chen and Dudhia 2001). The NOAH LSM simu-
lates soil moisture, soil temperature, skin temperature,
snowpack depth and water equivalent, canopy water
content, and the energy flux and water flux terms of the
surface energy balance and surface water balance. In its
MM5-coupled form NOAH has a diurnally dependent
Penman potential evaporation (Mahrt and Ek 1984), a
four-layer soil model (Mahrt and Pan 1984), a primitive
canopy model (Pan and Mahrt 1987), modestly com-
plex canopy resistance (Jacquemin and Noilhan 1990),
and a surface runoff scheme (Schaake et al. 1996). Ad-
ditionally, an irrigation scheme was implemented that
FIG. 6. Interannual standard deviation in absorbed short-
wave radiation (W m
2
), averaged over the month of May. Values
are calculated using NNRP incoming shortwave radiation and
AVHRR-derived surface albedo, 19812001. Note that maxima
exist in the semiarid pasture lands of the EP, Ustyurt Plateau
(east of the Caspian Sea), and Arabian Peninsula. Shading in-
dicates areas where the interannual standard deviation exceeds
15 W m
2
.
1A
UGUST 2007 Z A I TCHIK ET AL. 3931
realistically approximates flood irrigation practices of
the region (Zaitchik et al. 2005). The use of a coupled
landatmosphere modeling system was deemed neces-
sary in this study because of the tightly coupled nature
of interactions between surface fluxes, PBL dynamics,
and clouds (Betts 2004). For this reason, the results of
sensitivity studies performed with uncoupled LSMs can
be misleading in the analysis of physical process (Zhang
et al. 2001).
For all simulations the model was implemented at
27-km horizontal resolution and with 23 levels in the
vertical. Initial conditions and lateral boundary condi-
tions were drawn from the NNRP (Kalnay et al. 1996).
To capture the effects of interannual vegetation vari-
ability it was necessary to integrate satellite-derived
datasets for surface albedo and vegetation fraction. As
described in an earlier study (Zaitchik et al. 2005), op-
erational datasets currently available for MM5 provide
only climatological estimates for these properties.
Four 14-month MM5 simulations were performed.
First, control simulations were run for 1 November
199831 December 1999 and 1 November 200231 De-
cember 2003. In each case the first two months were
treated as spinup and results were recorded for the cal-
endar year. For these simulations surface albedo and
vegetation fraction were drawn from SPOT-Vegetation
data for the corresponding time period (see Section 2),
and they are hereafter referred to as 1999
V1999
and
2003
V2003
, where the subscript indicates the year of sat-
ellite data. It was necessary to prescribe albedo as well
as vegetation fraction because vegetation fields do not
inform the predicted surface albedo field in MM5
NOAH. Next, vegetation reversal simulations were
performed for the same time periods in which SPOT
albedo and vegetation fraction from the nondrought
year (2003) were input to MM5 simulations for the
drought year (1999) and vice versa. These simulations
are hereafter referred to as 1999
V2003
and 2003
V1999
.
It is important to note that the vegetation reversal
experiments include the reversal of vegetation fraction,
which is explicitly a property of vegetation, and surface
albedo, which may be influenced by a number of fac-
tors, notably soil moisture. It is the authors contention
that variability in the albedo field is associated prima-
rily with vegetation rather than soil moisture. The
SPOT images used to derive the surface albedo were
acquired in late morning under clear-sky conditions.
For the highly evaporative conditions that predominate
in the Middle East in late spring (the season of primary
interest), the soil surface is generally dry by this time of
day, minimizing the impact that residual soil moisture
can have on remotely sensed surface albedo. For this
reason it can be stated that vegetation reversal experi-
ments have the potential to capture vegetationclimate
feedbacks, but not soil moistureclimate feedbacks,
both of which may operate during a drought. If our
assumption is incorrect and albedo is under strong con-
trol of soil moisture, then soil moistureclimate feed-
backs related to albedo are included in the analysis,
while those related to the evaporation of soil water are
not. In either case, the inclusion of albedo in the veg-
etation reversal experiment leads to results that are dif-
ferent from vegetation reversal studies that have not
included albedo (e.g., Matsui et al. 2005).
c. Statistical analysis
For results presented as spatial averages, the signifi-
cance of differences between simulations was tested us-
ing a Students t test. To minimize the influence of spa-
tial autocorrelation, only a subset of grid points was
included in our statistical analysis, such that there was a
minimum distance of 108 km (four grid cells) separating
any two points included in the analysis. This resulted in
n 18 points for both the EP and northern Zagros
Plateau (NZ) subregions. The analysis cannot assess
FIG. 7. (A) CMAP (black lines) and MM5-simulated precipitation (gray lines), averaged monthly for the EP in 1999 (dashed lines)
and 2003 (solid lines). MM5-predicted total precipitation for (b) 1999 and (c) 2003, based on control simulations. The heavy contour
in (b) and (c) corresponds to 250 mm yr
1
, the theoretical limit for rain-fed agriculture in the region.
3932 JOURNAL OF CLIMATE VOLUME 20
the significance of results relative to the uncertainty in
atmospheric drivers, but it does establish the signifi-
cance of the differences relative to regional variability
in the MM5 simulations. All reported t tests are for
monthly averaged MM5 output.
6. Results of MM5 simulations
As expected, the MM5 returned drier conditions in
1999
V1999
than in 2003
V2003
. Precipitation results were
reasonably similar to the CMAP integrated observa-
tions (Fig. 7a), and precipitation was, in general, greater
in 2003 across the Euphrates Plain and the Zagros Pla-
teau (Figs. 7b and 7c). For the EP the difference in
precipitation was present only in the rainy and transi-
tional seasons, as JuneSeptember was almost com-
pletely dry in both years. Strong wintertime precipita-
tion in 2003 led to an extensive snowpack that persisted
through April in highland areas of Turkey and Iran.
For both years, snow retreated from most of the Tau-
rus and Zagros regions by May. May is an interesting
month for analysis because it lies in the transition be-
tween the rainy season of winter and spring and the
persistent aridity of summer. The intrusion of large
storm systems into the region is reduced relative to
NovemberApril, allowing for greater local influence
on the atmosphere, but significant precipitation is still
observed in most years, indicating a potential for moist
convection. For these reasons, May is a period during
which precipitation processes are expected to be sensi-
tive to a landatmosphere forcing involving persistent
spring vegetation or soil moisture.
From the perspective of landatmosphere interac-
tions, it is important to note that the 1999 drought led
to an increase in albedo only in the semiarid lowland
areas of the EP and, in the northeast of the study re-
gion, the Caspian steppe. In the Zagros foothills and
plateauwhere soil moisture is greater, soils are
darker, and vegetation is less sensitive to precipitation
droughtthere was no detectable difference in May
surface albedo between 1999 and 2003 (Fig. 8a; Table
1). In August a small difference between 1999 and 2003
albedo in the northern Zagros did develop (0.03, not
shown), but this difference was at all times smaller than
that in the EP, and it occurred at a time of year when
precipitation is strongly inhibited by large-scale circu-
lations (Rodwell and Hoskins 1996; Ziv et al. 2004),
limiting the potential for local feedbacks. This spatial
variability in albedo is important because of its impli-
cations for the surface radiation balance. If we consider
the two neighboring subregions of the EP and the NZ,
the MM5 simulations indicate that net radiation avail-
able at the surface (R
net
) was reduced in the EP ( p
0.0001) but not in the NZ (Fig. 8b). In fact, R
net
tended
to be slightly greater in 1999 for the Zagros, due to
clear-sky conditions and an increase in solar radiation
incident on the ground surface, though this tendency
was not significant ( p 0.359). This contrast in R
net
was associated with a contrast in ground temperature.
In the EP the warm synoptic conditions of 1999 were
offset by a decrease in local R
net
, such that the average
ground temperature in May was no higher in 1999 than
in 2003 (t test for difference: p 0.960). In the NZ, in
contrast, the increase in R
net
in 1999 combined with a
FIG. 8. Average difference for the month of May, 19992003, in (a) surface albedo and (b) net surface
radiation, according to MM5 simulations. Boxes EP and NZ indicate the Euphrates Plain and Northern
Zagros subregions described in the text and in Tables 1 and 2. In (b), regions of positive difference are
outlined by solid lines and regions of negative difference are outlined by dashed lines.
1A
UGUST 2007 Z A I TCHIK ET AL. 3933
substantial reduction in soil moisture to enhance large-
scale advective warming, producing radiative ground
temperatures that were 3.1°C warmer in May 1999 rela-
tive to May 2003 (Table 1; p 0.003). This MM5 result
is consistent with NNRP estimates of the 2-m air tem-
perature for the same period: according to the NNRP,
May 1999 was only 0.1°C warmer than May 2003 for the
EP, while it was 1.3°C warmer for the NZ.
The difference in drought impact on R
net
in the EP
versus the NZ also has an impact on turbulent heat flux
via the surface energy balance: R
net
H
E G,
where H is net surface sensible heat flux,
E is surface
latent heat flux, and G is conductive heat flux from the
surface to subsurface. The drought-related reduction in
R
net
in the EP for 1999 was associated with a decrease
in H of 6.1 W m
2
relative to 2003 ( p 0.0001), aver-
aged over the month of May, while the clear-sky-
associated increase in R
net
for the NZ was associated
with an increase in H of 32.4 W m
2
(p 0.0001).
Latent heat flux (
E), meanwhile, was reduced in 1999
for both the EP (by 9.6 W m
2
; p 0.0007) and the NZ
(by 29.7 W m
2
; p 0.0001) due to reduced soil mois-
ture. The effect was considerably larger in the NZ area
because soil moisture in the EP was quite low in May
even in the relatively wet year of 2003.
Vegetation reversal
Vegetation reversal experiments have the potential
to clarify the relative importance of synoptic forcing
and local vegetation properties during drought. The
reversals of the surface properties performed in this
experiment capture the effects of drought on albedo
and latent heat flux. Surface roughness effects were
not captured by these experiments, as roughness is not
a predicted field in the MM5NOAH simulations, and
the parameter was left at land cover defaults for all
simulations. Latent heat flux effects could arise as a
result of greater access to soil moisture, which is a func-
tion of vegetation fraction in MM5NOAH (Chen
and Dudhia 2001), or as a secondary product of albedo
since an increase in net radiation at the surface would
be expected to increase the total turbulent energy
transfer. In application, however, vegetation reversal
had no significant effect on the May latent heat flux
in the EP (Table 2; p 0.960). By this time of the
year model soil moisture was depleted in the EP in
both 1999 and 2003, leaving little possibility for vegeta-
tion to influence latent heat flux. What difference there
was in
E between 1999 and 2003 (Table 1) primarily
arose from direct soil evaporation after rain events, and
TABLE 1. Surface fluxes and PBL properties for MM5 control simulations. Values are averages for the month of May for albedo (
);
near-surface soil moisture (SM
010cm
); incoming and reflected solar radiation at the surface (S
, S
); surface-incoming and surface-
emitted longwave radiation (L
, L
); net surface radiation (R
net
); ground temperature (T
ground
); sensible heat flux (H ); latent heat flux
(
E ); conductive heat flux to the subsurface (G); depth and temperature of the planetary boundary layer (Depth
PBL
, T
PBL
); turbulent,
radiative, and total surface heat fluxes (Q
t
, Q
R
, Q
Total
); local contribution to the moist static energy density of the PBL (MSE
local
);
and precipitation.
Variable Unit
Euphrates Plain Northern Zagros
1999 2003 19992003 1999 2003 19992003
0.31 0.25 0.06 0.19 0.19 0.00
SM
010cm
m
3
m
3
0.132 0.160 0.028 0.148 0.215 0.067
S
Wm
2
344.7 330.5 14.2 331.8 301.9 29.9
S
Wm
2
106.9 82.6 24.2 63.0 57.4 5.7
L
Wm
2
337.4 344.2 6.8 306.4 311.3 4.9
L
Wm
2
459.3 459.3 0.0 424.2 406.6 17.6
R
net
Wm
2
116.0 132.8 16.8 151.0 149.3 1.7
T
ground
°C 300.0 300.0 0.0 294.1 291.0 3.1
H Wm
2
99.5 105.6 6.1 114.3 81.9 32.4
E Wm
2
6.6 16.2 9.6 25.4 55.1 29.7
G Wm
2
6.4 7.4 1.0 8.0 9.9 1.9
15.08 6.52 8.56 4.50 1.49 3.01
Depth
PBL
m 1270 1397 127 1477 1207 270
Q
t
Wm
2
106.1 121.8 15.7 139.7 137.0 2.7
T
PBL
* °C 293.1 292.1 1.0 281.8 278.9 2.9
Q
R
Wm
2
142.9 137.2 5.7 103.0 99.1 3.9
Q
total
Wm
2
36.8 15.4 21.4 36.7 37.9 1.2
MSE
local
Wm
3
0.029 0.011 0.018 0.025 0.031 0.007
Precip mm 0.6 10.1 9.5 4.6 20.3 15.7
*T
PBL
is taken at a height near the top of the average PBL: 850 hPa for EP and 700 hPa for NZ.
3934 JOURNAL OF CLIMATE VOLUME 20
this flux was not significantly affected by vegetation
status.
In the NZ, modeled
E for May was substantial in
both 1999 and 2003. In May 2003, in fact, soil moisture
was so great that actual evapotranspiration (AET) fre-
quently approached the atmospherically limited rate of
potential evapotranspiration (PET). Under such moist
conditions, the access to deeper reserves of soil mois-
ture afforded by healthy vegetation does not have a
substantial impact on
E. In 1999, when there was little
precipitation in May, the soil surface dried out and
E
was reduced by more than 50% (Table 1). Under these
drier conditions AET was well below PET due to mois-
ture limitation. Vegetation access to subsurface soil
moisture thus has the potential to influence the parti-
tioning of surface energy. In 1999, the average MEP
E
was reduced by 5.6 W m
2
in the simulation with
drought vegetation (1999
V1999
) relative to that with
nondrought vegetation (1999
V2003
), though the differ-
ence was not statistically significant (p 0.140).
Vegetation reversal had a substantial impact on ra-
diation fluxes in the EP for both 1999 and 2003 simu-
lations. In both years, drought vegetation was associ-
ated with increased albedo and reduced R
net
(p
0.0001). Reduced R
net
led to lower ground temperature
(significant for 2003, p
2003
0.050; nonsignificant for
1999, p
1999
0.109), reduced sensible heat flux ( p
0.0001), and a shallower PBL (p 0.0001) (Figs. 9ac).
Drought vegetation also led to a reduction in total tur-
bulent heat flux from the surface (Q
t
: Table 2) and an
elevated lifting condensation level (LCL: Fig. 9d), the
implications of which will be discussed in the following
section. In the NZ, where the 1999 drought did not
affect the May surface albedo, the vegetation reversal
experiment had no significant impact on radiation
fluxes in either year (R
net
p
1999
0.140, p
2003
0.597).
Ground temperature, sensible heat flux, and the depth
of the PBL were also unaffected (p 0.2 for all analy-
ses). As in the interannual comparison, the vegetation
reversal experiments did yield some effects on albedo
and R
net
for the NZ in July and August. These effects
were small relative to those in the EP in May and, as
mentioned above, they arise during a season when a
landatmosphere forcing on precipitation is unlikely to
occur.
7. Feedbacks on cloudiness and precipitation
Two hypotheses for a drought feedback on precipi-
tation are considered: a mechanism associated with re-
duced R
net
(Eltahir 1998) and a mechanism involving
deepening and drying of the planetary boundary layer
(Betts and Ball 1998). These hypotheses have been re-
viewed and compared in detail by Small and Kurc
(2003) and Zaitchik et al. (2006). Briefly, in the mecha-
nism proposed by Eltahir (1998) a drought-related in-
crease in albedo and decrease in soil moisture leads to
a reduction in R
net
and an associated reduction in total
TABLE 2. Impact of vegetation reversal on MM5 simulations. Values reported are differences that indicate the impact of drought
vegetation (1999
V1999
, 2003
V1999
) as opposed to healthy vegetation (1999
V2003
, 2003
V2003
) in each simulation year. Symbols are defined
in Table 1.
Variable Unit
Euphrates Plain Northern Zagros
1999
V1999
1999
V2003
2003
V1999
2003
V2003
1999
V1999
1999
V2003
2003
V1999
2003
V2003
0.06 0.06 0.00 0.00
SM
010cm
m
3
m
3
0.002 0.002 0.004 0.004
S
Wm
2
1.1 3.5 4.3 4.1
S
Wm
2
21.0 20.9 0.8 0.8
L
Wm
2
-2.9 4.1 2.8 1.4
L
Wm
2
6.2 6.7 0.0 0.0
R
net
Wm
2
16.6 14.8 0.7 1.9
T
ground
°C 1.0 1.1 0.0 0.0
H Wm
2
17.0 15.7 5.3 2.5
E Wm
2
0.3 0.3 5.6 0.6
G Wm
2
0.2 1.1 0.0 0.3
Depth
PBL
m 165 156 54
Q
t
Wm
2
17.3 16.0 0.3 1.9
T
PBL
* °C 0.5 0.5 0.1 0.1
Q
R
Wm
2
0.8 1.2 0.5 0.5
Q
total
Wm
2
18.1 17.2 0.2 2.4
MSE
local
Wm
3
0.016 0.015 0.000 0.002
Precip mm 0.0 0.9 0.4 0.8
* T
PBL
is taken at a height near the top of the average PBL: 850 hPa for EP and 700 hPa for NZ.
1A
UGUST 2007 Z A I TCHIK ET AL. 3935
turbulent heat flux from the surface (Q
t
H
E).
This means that the local land surface contributes less
to the moist static energy (MSE) of the PBL during
drought than it would under normal conditions. The
density of MSE in the PBL is directly associated with
conditional instability (Zawadzki et al. 1981), so a re-
duction in the local contribution to MSE leads to re-
duced potential for convective precipitation. The Betts
and Ball (1998) hypothesis focuses on total local heat-
ing of the PBL. Because latent heat flux is not released
locallycondensation might occur some distance
downwindit is excluded from consideration. Instead,
the hypothesis states that an increase in H, possibly
accompanied by an increase in local radiative heating
(Q
R
), causes a deepening of the PBL due to more vig-
orous mixing and entrainment at the top of the bound-
ary layer. The combination of deepening and entrain-
ment of dry, low MSE air from the free troposphere
leads to reduced MSE density (MSE) and a higher
lifting condensation level (LCL), thus reducing the po-
tential for precipitation.
Comparing 1999 to 2003, the average R
net
for May in
the EP was 16.8 W m
2
less during the drought year of
1999, resulting in a 15.7 W m
2
decrease in Q
t
(Table
1). Additionally, a slightly cooler ground temperature
in 1999 was associated with a 5.7 W m
2
reduction in
the net radiative exchange (Q
R
) between the surface
and the lower atmosphere. Here Q
R
was calculated as a
simple radiative balance, Q
R
⫽␧
a
L
⫺␧
s
L
, where
a
and
s
are the emissivity of the lower atmosphere and
surface, and longwave upwelling (L
) and downwelling
(L
) terms are solved based on ground temperature
and air temperature near the top of the PBL, respec-
tively (Liou 2002; Zaitchik et al. 2006). The change in
Q
t
and Q
R
sum to a reduction of 22.5 W m
2
in the total
local energy transfer between the surface and the lower
atmosphere (Q
Total
Q
t
Q
R
). A forcing of this mag-
nitude is on the order of forcings that have triggered
precipitation feedbacks in previous modeling studies
(e.g., Schär et al. 1999), though the prevailing aridity of
the Middle East may limit the potential for feedback in
this region.
In the NZ, in contrast, the 1999 drought was associ-
ated with very slight increases in average R
net
(1.7 W
m
2
) and Q
t
(2.7 W m
2
) for the month of May. The
small change in Q
t
is the result of large offsetting
changes in H (32.4 W m
2
in the drought year) and
E (29.7 W m
2
). The minor increase in Q
t
was offset
by a decrease in Q
R
, and the difference in Q
Total
be-
tween 1999 and 2003 (1.2Wm
2
) was far too small to
trigger any forcing on precipitation via the Eltahir
(1998) mechanism. The MM5 results for the NZ are
consistent with the Betts and Ball (1998) feedback hy-
pothesis, however, as otal local heating (H Q
R
) was
enhanced by 28.5 W m
2
in 1999 relative to 2003 and
the average depth of the PBL was increased by 87 m.
This leads to a deeper, drier PBL, with reduced poten-
tial for moist convection.
Vegetation reversal
Cloud thickness in MM5 can be estimated by sum-
ming the column-integrated fields of all condensed wa-
ter species (ice, snow, rainwater, and cloud water). Av-
eraged over a month, this cloudiness calculation is a
proxy for total cloudiness, sensitive to changes in either
cloud frequency or cloud thickness. The spatial pattern
of the MM5 cloudiness compares relatively well with
the MODIS-derived monthly cloud fraction (e.g., Fig.
10).
Comparing simulations 1999
V2003
and 1999
V1999
(Fig. 11a), one finds that cloudiness was greater over
the northern Zagros and east of the Caspian in the
1999
V2003
simulation. There was essentially no differ-
ence in cloudiness over the EP. In 2003, when the back-
ground atmospheric humidity and local evaporation
were both higher than in 1999 (Table 2), the presence of
nondrought vegetation (2003
V2003
2003
V1999
) results
in increased cloudiness over broad areas of the EP,
Saudi Arabia, Turkey, the Zagros Plateau, and east of
the Caspian Sea (Fig. 11b). For the EP this discrepancy
between the 1999 and 2003 vegetation reversal ex-
FIG. 9. Results of MM5 vegetation reversal experiment for 1999,
expressed as the difference 1999
V1999
1999
V2003
: (a) ground
temperature (°C), (b) sensible heat flux (W m
2
), (c) PBL depth
(m), and (d) lifting condensation level (m). Contour intervals are
0.4 (a), 10.0 (b), (d), and 50.0 (c), centered on zero, with negative
contours dashed.
3936 JOURNAL OF CLIMATE VOLUME 20
periments is most pronounced in May (Fig. 11c). This
supports the suggestion that May is a key month for
landatmosphere forcings on hydrometeorology (sec-
tion 5). Over the NZ differences in cloudiness are
present from April through August in both 1999 and
2003 vegetation reversal experiments, with increased
cloudiness observed in simulations with healthy vegeta-
tion (1999
V2003
and 2003
V2003
).
The cloudiness result for the EP emphasizes the im-
portance of abiotic factorsprimarily soil moisture and
mesoscale atmospheric conditionsin determining
whether a vegetation forcing will trigger a feedback on
the atmosphere. In the wetter, convectively active con-
ditions of 2003 the presence of a healthy vegetation
cover had the potential to enhance the production of
clouds relative to a simulation with drought vegetation.
The dry, convectively stable background conditions of
1999, meanwhile, were relatively resistant to the intro-
duction of an artificially healthy vegetation cover
(1999
V2003
), and cloudiness effects were not realized in
the EP.
It is difficult to assess the significance of the cloudi-
ness and precipitation results relative to background
variability without performing an ensemble of simula-
tions, which was not possible in the presented research.
Nonetheless, differences in cloudiness in the vegetation
reversal experiments suggest that a vegetation-induced
feedback is operating. This evidence of feedback in the
MM5 can be attributed to the reversal of vegetation
fraction and albedo fields because they were the only
variables manipulated in the experiment and because
no perturbations were applied to the forcing data.
For the EP, healthy vegetation cover is associated
with reduced albedo, enhanced sensible heat flux, and
FIG. 11. Column-integrated condensed water content (g m
2
): (a) 1999
V1999
1999
V2003
, (b) 2003
V1999
2003
V2003
, average
values for the month of May. Contour interval is8gm
2
, centered on zero, with negative contours dashed. (c) Time series
of monthly averages for 1999
V1999
1999
V2003
(solid line) and 2003
V1999
2003
V2003
(dashed line), averaged over
the EP.
FIG. 10. Cloud cover for May 2003: (a) MODIS monthly average cloud fraction, composited
at 1° from MODIS Terra and MODIS Aqua data; (b) MM5 2003
V2003
column-integrated
condensed water content (g m
2
), averaged for the month of May and smoothed for com-
parison with MODIS.
1A
UGUST 2007 Z A I TCHIK ET AL. 3937
increased MSE density within the PBL (Table 2). This
is consistent with the R
net
drought feedback proposed
by Eltahir (1998) and with results of the interannual
model comparison described above. In the NZ the veg-
etation reversal experiments had little impact on sur-
face radiation and total turbulent heat flux. This sug-
gests that vegetation cover on its own does not signifi-
cantly influence landatmosphere interactions during
drought. Instead, soil moisture may play a more domi-
nant role, through its effects on soil surface properties
and its relevance to evapotranspiration. This being the
case, differences in NZ cloud cover obtained in vegeta-
tion reversal simulations (Figs. 11a and 11b) can be
explained in two ways. First, the NZ lies downwind of
the EP, so any hydrometeorological forcing in the EP
has the potential to impact cloud formation in the NZ.
Second, for mountainous areas such as the NZ, the veg-
etation reversal experiments include some soil moisture
effects. This is because portions of the NZ were snow
covered through March 2003, while the area was nearly
snow-free in March 1999. This snow cover discrepancy
affected the SPOT-derived surface albedo, with the re-
sult that simulations using the 2003 albedo (i.e.,
1999
V2003
and 2003
V2003
) had less early spring evapora-
tion and thus greater residual soil moisture in late
spring than the simulations that used the 1999 albedo.
Finally, it should be noted that secondary effects of
vegetation-induced cloud coverreduced incident
shortwave radiation and enhanced downwelling long-
wave radiationwere small and offsetting in both re-
gions for both 1999 and 2003 (Table 2).
The distinction between the EP and the NZ is not
particularly surprising. In the EP surface moisture is
extremely limited in late spring, and evapotranspiration
is limited as well. The greatest difference between a
drought year and a wet year, then, is that in a wet year
the surface albedo is reduced due to greater vegetation
cover over bright, dry soils. In the NZ, as indicated in
satellite analysis (Fig. 6), interannual variability in al-
bedo is relatively small. Instead, a year of below-
average precipitation results in substantial reductions in
soil moisture and evapotranspiration. These soil-
moisture-induced changes in the surface energy balance
form the basis for variability in landatmosphere inter-
actions.
8. Conclusions
In this study it was found that interannual fluctua-
tions in climate lead to considerable variability in veg-
etation for much of the Middle East. The greatest vari-
ability was found in a sensitive transitional climate zone
that includes much of the Fertile Crescent and its neigh-
boring rangelands. In this area vegetation intensity is
strongly correlated with surface albedo both during and
after the prime growing season. An externally imposed
drought, then, leads to reduced vegetation, increased
albedo, and the potential for feedbacks associated with
changed land surface conditions.
During the drought of 1999, the impact of drought on
vegetation, albedo, and soil moisture led to conditions
consistent with surface feedbacks on precipitation. In
the Euphrates Plain (EP), increased albedo led to re-
duced local heat flux, producing a shallow, stable plan-
etary boundary layer (PBL) with low conditional insta-
bility (Fig. 12). In the neighboring northern Zagros re-
gion (NZ), drought led to reduced soil moisture and
increased sensible heat flux, causing a deep, dry PBL to
develop. This condition is associated with enhanced en-
trainment of dry air at the top of the PBL and a reduc-
tion in conditional instability (Fig. 12). In vegetation
reversal sensitivity experiments it was found that the
drought-related feedback tendency in the EP relied
heavily on vegetation status and albedo, while that in
the NZ appeared to be more strongly associated with
near-surface soil moisture.
The EP is substantially more arid than the NZ. Dif-
ferences in drought impacts and feedbacks between the
two subregions are analogous to differences observed
during drought in predominantly arid zones versus
more humid regions. In dry areas, vegetation cover is
sparse and soils are typically bright. Vegetation drought
therefore causes an increase in albedo and a reduced
energy environment (Charney et al. 1977; Eltahir 1998).
In more humid regions, drought is not necessarily as-
sociated with an increase in albedo; senescent vegeta-
tion tends to be brighter than live vegetation, and dry
FIG. 12. Surface processes relevant to droughtprecipitation
forcings in the EP and NZ (schematic): symbols as defined in
Table 1. Solid gray line represents the height of the PBL under
nondrought conditions and the dashed line represents the height
of the PBL during drought. The black line is the land surface.
3938 JOURNAL OF CLIMATE VOLUME 20
soils are brighter than wet soils (Eltahir 1998), but these
effects can be mitigated by resilient vegetation and off-
set by increased exposure of soils that are relatively
dark even when dry (Zaitchik et al. 2006). Under these
circumstances feedbacks related to evapotranspiration
are expected to dominate, as drought transforms a nor-
mally moisture-rich landscape into a moisture-limited
environment (e.g., Heck et al. 1999; Sud et al. 2003).
In the Middle East, climate variability is largely a
product of external factors, and drought is a common
occurrence for both the Euphrates Plain and the more
humidbut still drought-proneZagros Plateau. In a
region that contains steep precipitation gradients and
large areas of marginal rain-fed agriculture, any local
processes that enhance or mitigate drought are of in-
terest. These processes can include local forcings on air
temperature, feedbacks on vapor pressure and cloudi-
ness, and modification of precipitation events. The
present study revealed strong evidence for a local in-
fluence on temperature, water vapor, and cloudiness
during moderate drought, with inconclusive results on
precipitation. Further studies of Middle East drought,
including analysis of historically extreme or persistent
droughts, will further our understanding of landatmo-
sphere interactions in this environmentally sensitive re-
gion.
Acknowledgments. This study was carried out with
the financial support of NASA Grant NNG05GB36G
and with NCAR computer resources made available
under NSF Grant ATM-0112354. Preprocessed
AVHRR data were provided by Prasad Thenkabail and
Praveen Noojipady of the International Water Manage-
ment Institute (Colombo, Sri Lanka).
REFERENCES
Archer, S., D. S. Schimel, and E. A. Holland, 1995: Mechanisms of
shrubland expansion: Land-use, climate or CO
2
. Climatic
Change, 29, 9199.
Baldocchi, D. D., L. K. Xu, and N. Kiang, 2004: How plant func-
tional-type, weather, seasonal drought, and soil physical
properties alter water and energy fluxes of an oak-grass sa-
vanna and an annual grassland. Agric. For. Meteor., 123, 13
39.
Betts, A. K., 2004: Understanding hydrometeorology using global
models. Bull. Amer. Meteor. Soc., 85, 16731688.
——, and J. H. Ball, 1998: FIFE surface climate and site-average
dataset 198789. J. Atmos. Sci., 55, 10911108.
Bright, D. R., and S. L. Mullen, 2002: The sensitivity of the nu-
merical simulation of the Southwest monsoon boundary layer
to the choice of PBL turbulence parameterization in MM5.
Wea. Forecasting, 17, 99114.
Brunsell, N. A., 2006: Characterization of land-surface precipita-
tion feedback regimes with remote sensing. Remote Sens. En-
viron., 100, 200211.
Charney, J. G., 1975: Dynamics of deserts and drought in Sahel.
Quart. J. Roy. Meteor. Soc., 101, 193202.
——, W. J. Quirk, S. H. Chow, and J. Kornfield, 1977: Compara-
tive study of effects of albedo change on drought in semi-arid
regions. J. Atmos. Sci., 34, 13661385.
Chen, F., and K. Mitchell, 1999: Using GEWEX/ISLSCP forcing
data to simulate global soil moisture fields and hydrological
cycle for 19871988. J. Meteor. Soc. Japan, 77, 116.
——, and J. Dudhia, 2001: Coupling an advanced land surface-
hydrology model with the Penn StateNCAR MM5 modeling
system. Part II: Preliminary model validation. Mon. Wea.
Rev., 129, 587604.
Courel, M. F., R. S. Kandel, and S. I. Rasool, 1984: Surface albedo
and the Sahel drought. Nature, 307, 528531.
DallOlmo, G., and A. Karnieli, 2002: Monitoring phenological
cycles of desert ecosystems using NDVI and LST data de-
rived from NOAA-AVHRR imagery. Int. J. Remote Sens.,
23, 40554071.
deMenocal, P. B., 2001: Cultural responses to climate change dur-
ing the Late Holocene. Science, 292, 667673.
DePauw, E., 2004: Drought early warning systems for the Near
East. Challenges and Strategies for Dryland Agriculture, S. C.
Rao and J. Ryan, Eds., American Society of Agronomy and
Crop Science Society of America, 93111.
Dudhia, J., 1993: A nonhydrostatic version of the Penn State
NCAR mesoscale model: Validation tests and simulation of
an Atlantic cyclone and cold front. Mon. Wea. Rev., 121,
14931513.
Eltahir, E. A. B., 1998: A soil moisture rainfall feedback mecha-
nism 1. Theory and observations. Water Resour. Res., 34, 765
776.
Evans, J. P., and R. Geerken, 2004: Discrimination between cli-
mate and human-induced dryland degradation. J. Arid Envi-
ron., 57, 535554.
——, and R. B. Smith, 2005: Classifying precipitation events in the
Fertile Crescent. Preprints, 16th Conf. on Climate Variability
and Change, San Diego, CA, Amer. Meteor. Soc., CD-ROM,
9.4.
——, and R. Geerken, 2006: Classifying rangeland vegetation type
and coverage using a Fourier component based similarity
measure. Int. J. Remote Sens., 105, 18.
FAO, 1997: Irrigation in the Near East region in figures. Water
Reports 9, U.N. Food and Agriculture Organization, 292 pp.
[Available online at http://www.fao.org/docrep/w4356e/
w4356e00.htm.]
Geerken, R., and M. Ilaiwi, 2004: Assessment of rangeland deg-
radation and development of a strategy for rehabilitation.
Remote Sens. Environ., 90, 490504.
——, N. Batikha, D. Celis, and E. Depauw, 2005a: Differentiation
of rangeland vegetation and assessment of its status: Field
investigations and MODIS and SPOT VEGETATION data
analyses. Int. J. Remote Sens., 26, 44994526.
——, B. Zaitchik, and J. P. Evans, 2005b: Classifying rangeland
vegetation type and coverage from NDVI time series using
Fourier filtered cycle similarity. Int. J. Remote Sens., 26,
55355554.
Grell, G. A., J. Dudhia, and D. R. Stauffer, 1994: A description of
the fifth-generation Penn State/NCAR Mesoscale Model
(MM5). NCAR Tech. Note NCAR/TN-398STR, Boulder,
CO, 117 pp.
Gutman, G., and A. Ignatov, 1998: The derivation of the green
vegetation fraction from NOAA/AVHRR data for use in
1A
UGUST 2007 Z A I TCHIK ET AL. 3939
numerical weather prediction models. Int. J. Remote Sens.,
19, 15331543.
Heck, P., D. Luthi, and C. Schär, 1999: The influence of vegeta-
tion on the summertime evolution of European soil moisture.
Phys. Chem. Earth, 24B, 609614.
Hole, F., 1994: Environmental instabilities and urban origins.
Chiefdoms and Early States in the Near East: the Organiza-
tional Dynamics of Complexity, Prehistory Press, 121151.
Hong, S. Y., and H. L. Pan, 1996: Nonlocal boundary layer verti-
cal diffusion in a medium-range forecast model. Mon. Wea.
Rev., 124, 23222339.
Jacquemin, B., and J. Noilhan, 1990: Sensitivity study and valida-
tion of a land surface parameterization using the HAPEX-
MOBILHY data set. Bound.-Layer Meteor., 52, 93134.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Re-
analysis Project. Bull. Amer. Meteor. Soc., 77, 437471.
Koster, R. D., and Coauthors, 2004: Regions of strong coupling
between soil moisture and precipitation. Science, 305, 1138
1140.
Kremer, R. G., E. R. Hunt, S. W. Running, and J. C. Coughlan,
1996: Simulating vegetational and hydrologic responses to
natural climatic variation and GCM-predicted climate change
in a semi-arid ecosystem in Washington, USA. J. Arid Envi-
ron., 33, 2338.
Lane, D. R., D. P. Coffin, and W. K. Lauenroth, 1998: Effects of
soil texture and precipitation on above-ground net primary
productivity and vegetation structure across the central grass-
land region of the United States. J. Veg. Sci., 9, 239250.
Laval, K., and L. Picon, 1986: Effect of a change of the surface
albedo of the Sahel on climate. J. Atmos. Sci., 43, 24182429.
LeHouerou, H. N., 1996: Climate change, drought and desertifi-
cation. J. Arid Environ., 34, 133185.
Liang, S. L., C. J. Shuey, A. L. Russ, H. L. Fang, M. Z. Chen, C. L.
Walthall, C. S. T. Daughtry, and R. Hunt, 2002: Narrowband
to broadband conversions of land surface albedo: II. Valida-
tion. Remote Sens. Environ., 84, 2541.
Liou, K. N., 2002: An Introduction to Atmospheric Radiation.2d
ed. International Geophysics Series, Vol. 84, Academic Press,
583 pp.
Los, S. O., 1993: Calibration adjustment of the NOAA AVHRR
normalized difference vegetation index without recourse to
component channel-1 and channel-2 data. Int. J. Remote
Sens., 14, 19071917.
Mahrt, L., and M. Ek, 1984: The influence of atmospheric stability
on potential evaporation. J. Climate Appl. Meteor., 23, 222
234.
——, and H. L. Pan, 1984: A two-layer model of soil hydrology.
Bound.-Layer Meteor., 29, 120.
Maisongrande, P., B. Duchemin, and G. Dedieu, 2004: VEGETA-
TION/SPOT: An operational mission for the Earth monitor-
ing; presentation of new standard products. Int. J. Remote
Sens., 25, 914.
Matsui, T., V. Lakshmi, and E. E. Small, 2005: The effects of
satellite-derived vegetation cover variability on simulated
landatmosphere interactions in the NAMS. J. Climate, 18,
2140.
Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A.
Clough, 1997: Radiative transfer for inhomogeneous atmo-
sphere: RRTM, a validated correlated-k model for the long-
wave. J. Geophys. Res., 102, 16 66316 682.
Myhre, G., E. J. Highwood, K. P. Shine, and F. Stordal, 1998: New
estimates of radiative forcing due to well mixed greenhouse
gases. Geophys. Res. Lett., 25, 27152718.
Nielsen, T. T., and H. K. Adriansen, 2005: Government policies
and land degradation in the Middle East. Land Degradation
Development, 16, 151161.
Pan, H. L., and L. Mahrt, 1987: Interaction between soil hydrol-
ogy and boundary-layer development. Bound.-Layer Meteor.,
38, 185202.
Reisner, J., R. J. Rasmussen, and R. T. Bruintjes, 1998: Explicit
forecasting of supercooled liquid water in winter storms using
the MM5 mesoscale model. Quart. J. Roy. Meteor. Soc., 124B,
10711107.
Rodwell, M. J., and B. J. Hoskins, 1996: Monsoons and the dy-
namics of deserts. Quart. J. Roy. Meteor. Soc., 122, 1385
1404.
Schaake, J. C., V. I. Koren, Q. Y. Duan, K. Mitchell, and F. Chen,
1996: A simple water balance model (SWB) for estimating
runoff at different spatial and temporal scales. J. Geophys.
Res., 101, 74617475.
Schär, C., D. Luthi, U. Beyerle, and E. Heise, 1999: The soil
precipitation feedback: A process study with a regional cli-
mate model. J. Climate, 12, 722741.
——, P. L. Vidale, D. Luthi, C. Frei, C. Haberli, M. A. Liniger,
and C. Appenzeller, 2004: The role of increasing temperature
variability in European summer heatwaves. Nature, 427, 332
336.
Schmidt, H., and A. Karnieli, 2000: Remote sensing of the sea-
sonal variability of vegetation in a semi-arid environment. J.
Arid Environ., 45, 4359.
Shukla, J., and Y. Mintz, 1982: Influence of land-surface evapo-
transpiration on the earths climate. Science, 215, 14981501.
Small, E. E., and S. A. Kurc, 2003: Tight coupling between soil
moisture and the surface radiation budget in semiarid envi-
ronments: Implications for land-atmosphere interactions.
Water Resour. Res., 39, 12781291.
Sud, Y. C., and M. Fennessy, 1982: A study of the influence of
surface albedo on July circulation in semi-arid regions using
the GLAS GCM. J. Climatol., 2, 105125.
——, and M. J. Fennessy, 1984: Influence of evaporation in semi-
arid regions on the July circulationA numerical study. J.
Climatol., 4, 383398.
——, and W. E. Smith, 1985: The influence of surface-roughness
of deserts on the July circulation (a numerical study).
Bound.-Layer Meteor., 33, 1549.
——, D. M. Mocko, K. M. Lau, and R. Atlas, 2003: Simulating the
midwestern U.S. drought of 1988 with a GCM. J. Climate, 16,
39463965.
Vanacker, V., M. Linderman, F. Lupo, S. Flasse, and E. Lambin,
2005: Impact of short-term rainfall fluctuation on interannual
land cover change in sub-Saharan Africa. Global Ecol. Bio-
geogr., 14, 123135.
Weiss, E., S. E. Marsh, and E. S. Pfirman, 2001: Application of
NOAA-AVHRR NDVI time-series data to assess changes in
Saudi Arabias rangelands. Int. J. Remote Sens., 22, 1005
1027.
Weiss, H., and R. S. Bradley, 2001: What drives societal collapse?
Science, 291, 609610.
Weiss, J. L., D. S. Gutzler, J. E. A. Coonrod, and C. N. Dahm,
2004: Long-term vegetation monitoring with NDVI in a di-
verse semi-arid setting, central New Mexico, USA. J. Arid
Environ., 58, 249272.
Xie, P. P., and P. A. Arkin, 1996: Analyses of global monthly
precipitation using gauge observations, satellite estimates,
and numerical model predictions. J. Climate, 9, 840858.
Zaitchik, B. F., J. Evans, and R. B. Smith, 2005: MODIS-derived
3940 JOURNAL OF CLIMATE VOLUME 20
boundary conditions for a mesoscale climate model: Appli-
cation to irrigated agriculture in the Euphrates basin. Mon.
Wea. Rev., 133, 17271743.
——, A. K. Macalady, L. R. Bonneau, and R. B. Smith, 2006: Eu-
ropes 2003 heat wave: A satellite view of impacts and land
atmosphere feedbacks. Int. J. Climatol., 26, 743769.
Zawadzki, I., E. Torlaschi, and R. Sauvageau, 1981: The relation-
ship between mesoscale thermodynamic variables and con-
vective precipitation. J. Atmos. Sci., 38, 15351540.
Zhang, H., A. Henderson-Sellers, A. J. Pitman, J. L. McGregor,
C. E. Desborough, and J. J. Katzfey, 2001: Limited-area
model sensitivity to the complexity of representation of the
land surface energy balance. J. Climate, 14, 39653986.
Zheng, Y. Q., G. Yu, Y. F. Qian, M. Q. Miao, X. M. Zeng, and
H. Q. Liu, 2002: Simulations of regional climatic effects of
vegetation change in China. Quart. J. Roy. Meteor. Soc., 128,
20892114.
Ziv, B., H. Saaroni, and P. Alpert, 2004: The factors governing the
summer regime of the eastern Mediterranean. Int. J. Clima-
tol., 24, 18591871.
1A
UGUST 2007 Z A I TCHIK ET AL. 3941
... This anomaly is modest in magnitude, but it represents a substantial reduction for this generally arid and semi-arid region ( Figure S1). As one might expect in a climatologically dry region (Zaitchik et al., 2007), this reduction in vegetation and generally low soil moisture conditions (Figure 2b) are associated with an increase in albedo ( Figure 2c). Elevated albedo reduces the amount of solar radiation absorbed at the surface and is associated with a negative forcing on near-surface air temperatures (Zaitchik et al., 2007). ...
... As one might expect in a climatologically dry region (Zaitchik et al., 2007), this reduction in vegetation and generally low soil moisture conditions (Figure 2b) are associated with an increase in albedo ( Figure 2c). Elevated albedo reduces the amount of solar radiation absorbed at the surface and is associated with a negative forcing on near-surface air temperatures (Zaitchik et al., 2007). At the same time, satellite-derived estimates show reduced rates of evapotranspiration over a substantial fraction of the drought-affected area (Figure 2d), including in forested areas of Arizona in the core of the dry-hot extreme (Figure 2e), which is typically associated with a shift from latent to sensible heat flux, exerting a positive forcing on air temperature. ...
... The Southwest U.S. heatwave of June 2021 was concentrated in a semi-arid and arid region with low vegetation cover and typically bright, low organic content soils. In this setting, a drought that causes vegetation to die back will expose a high albedo dry soil surface, and the reflection of solar radiation by this bright surface will reduce net radiation at the surface and associated turbulent heat flux to the atmosphere (Eltahir, 1998;Zaitchik et al., 2007). The situation can be quite different in transitional climate zones of strong land-atmosphere coupling where drought can cause dramatic reductions in evapotranspiration (Koster et al., 2004;Seneviratne et al., 2010), while the impact on vegetation does not always lead to a dramatic increase in albedoeither vegetation is dense enough to maintain coverage over soils, or the soils that are revealed by vegetation die-back are high organic matter soils that are relatively dark even when dry. ...
Article
Full-text available
In June of 2021 the southwest United States experienced a record-breaking heatwave. This heatwave came at a time when the region was in severe drought. As drought alters the surface energy budget in ways that affect lower atmosphere temperature and circulations, it is possible that the combined drought-heat event was a cascading climate hazard, in which preexisting drought exacerbated the heatwave. We apply satellite observation and numerical experiments with the Weather Research and Forecasting (WRF) model to test for land-atmosphere feedbacks during the heatwave consistent with drought influence. We find a modest positive drought-heat effect, as WRF simulations that include the drought have marginally higher air temperatures than those that exclude the initial drought conditions, with more substantial effects in wetter, forested areas. Evidence of drought-heat-drought coupled feedbacks was similarly modest in our simulations, as accounting for drought preconditioning led to a small reduction in simulated precipitation in the region.
... According to World Atlas, ME consists of 17 countries (Iran, Iraq, Bahrain, Egypt, Cyprus, Jordan, Israel, Kuwait, Oman, Lebanon, Palestine, Saudi Arabia, Qatar, Turkey, The Syrian Arab Republic, Yemen, and the United Arab Emirates) [75]. The ME is a predominantly semiarid region that contains a strong north to south precipitation gradient [76]. Annual precipitation may reach more than 1000 mm in humid regions of Turkey and Transcaucasia, while the deserts south of the Euphrates River may receive up to 100 mm/year [76]. ...
... The ME is a predominantly semiarid region that contains a strong north to south precipitation gradient [76]. Annual precipitation may reach more than 1000 mm in humid regions of Turkey and Transcaucasia, while the deserts south of the Euphrates River may receive up to 100 mm/year [76]. According to the recent GlobeLand30, cultivated land, grassland, shrubland, and forest are occupying 11.11%, 9.47%, 2.69%, and 2.69% of ME land area in 2020 (Figure 1b) [77]. ...
Article
Full-text available
Global change impacts including climate change, increased CO2 and nitrogen deposition can be determined through a more precise characterisation of Land Surface Phenology (LSP) parameters. In addition, accurate estimation of LSP dates is being increasingly used in applications such as mapping vegetation types, yield forecasting, and irrigation management. However, there has not been any attempt to characterise Middle East vegetation phenology at the fine spatial resolution appropriate for such applications. Remote-sensing based approaches have proved to be a useful tool in such regions since access is restricted in some areas due to security issues and their inter-annual vegetation phenology parameters vary considerably because of high uncertainty in rainfall. This study aims to establish for the first time a comprehensive characterisation of the vegetation phenological characteristics of the major vegetation types in the Middle East at a fine spatial resolution of 30 m using Landsat Normalized Difference Vegetation Index (NDVI) time series data over a temporal range of 20 years (2000–2020). Overall, a progressive pattern in phenophases was observed from low to high latitude. The earliest start of the season was concentrated in the central and east of the region associated mainly with grassland and cultivated land, while the significantly delayed end of the season was mainly distributed in northern Turkey and Iran corresponding to the forest, resulting in the prolonged length of the season in the study area. There was a significant positive correlation between LSP parameters and latitude, which indicates a delay in the start of the season of 4.83 days (R2 = 0.86, p < 0.001) and a delay in the end of the season of 6.54 days (R2 = 0.83, p < 0.001) per degree of latitude increase. In addition, we have discussed the advantages of fine resolution LSP parameters over the available coarse datasets and showed how such outputs can improve many applications in the region. This study shows the potential of Landsat data to quantify the LSP of major land cover types in heterogeneous landscapes of the Middle East which enhances our understanding of the spatial-temporal dynamics of vegetation dynamics in arid and semi-arid settings in the world.
... Here, the air mass came from the north (see the HYSPLIT air mass back trajectory in Fig. 3), passing over the eastern side of the Black Sea and over the region known as the Fertile Crescent (the approximate location of the ancient Fertile Crescent is shown in Fig. 3), which spans the northern regions of Israel, Syria, and Iraq. The Fertile Crescent is the historical origin of agriculture and animal herding that started around 12 000 years ago (Salamini et al., 2002), though in modern times it has suffered from long-term drying and is increasingly prone to drought (Kelley et al., 2015;Zittis et al., 2022), but it still remains an active location for agriculture and vegetation (Zaitchik et al., 2007). Given this unique air mass trajectory amongst the rest of the dataset, the high INP concentrations at warmer temperatures, and the "hump" in the INP spectrum that may indicate biological INP content as mentioned earlier, these data are hence discussed in further detail later in the article. ...
Article
Full-text available
While the atmosphere in the eastern Mediterranean is part of the dust belt, it encounters air masses from Europe, the Mediterranean Sea, and the Sahara and Arabian Desert that bring with them a whole host of potential dust and bioaerosol compositions and concentrations via long-range transport. The consequential changes in the populations of ice-nucleating particles (INPs), aerosols that influence weather and climate by the triggering of freezing in supercooled cloud water droplets, including in the convective cloud systems in the region, are not so well understood beyond the influence of desert dust storms in increasing INP concentrations. Here, we undertook an intensive INP measurement campaign in Israel to monitor changes in concentrations and activity from four major air masses, including the potential for activity from biological INPs. Our findings show that the INP activity in the region is likely dominated by the K-feldspar mineral content, with southwesterly air masses from the Sahara and easterly air masses from the Arabian Desert markedly increasing both aerosol and INP concentrations. Most intriguingly, a handful of air masses that passed over the Nile Delta and the northern Fertile Crescent, regions containing fertile agricultural soils and wetlands, brought high INP concentrations with strong indicators of biological activity. These results suggest that the Fertile Crescent could be a sporadic source of high-temperature biological ice-nucleating activity across the region that could periodically dominate the otherwise K-feldspar-controlled INP environment. We propose that these findings warrant further exploration in future studies in the region, which may be particularly pertinent given the ongoing desertification of the Fertile Crescent that could reveal further sources of dust and fertile soil-based INPs in the eastern Mediterranean region.
... Apart from these mountainous regions, a large part of Middle East is occupied by barren lands with little vegetation cover including the Fertile Crescent region and the Arabian Desert. These barren lands can serve as potential dust source regions (Notaro et al., 2015;Shohami et al., 2011;Zaitchik et al., 2007). One of the largest sand deserts in the world is the Rub al Khali located in the southern interior of the Arabian Peninsula. ...
Article
Full-text available
The assessment of Earth System Models (ESMs) in simulating long-term dust activities is a prerequisite for the applications of ESMs to the projection of future dust change. Middle East is the second largest dust source region in the world after North Africa, and our knowledge of the ESMs' ability in simulating regional dust cycle in this region is still limited. Here we examine the ability of the Chinese Academy of Sciences Earth System Model version 2 (CAS-ESM2) in simulating the spatiotemporal variations of dust activity over the Middle East region (30°E to 65°E, 10°N to 42°N). CAS-ESM2 simulations are compared with station observations, satellite observations, the Dust Constraints from joint Experimental-Modeling-Observational Analysis (DustCOMM) dataset, and the Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2). The results show that CAS-ESM2 captures the main dust sources extending from the south of the Persian Gulf to the Arabian Peninsula, Iraq, and Sudan. CAS-ESM2 also simulates similar patterns of dust optical depth (DOD) with referential data, although there is some uncertainty in DOD observation. Further analysis shows that the model is able to reproduce the temporal variations of dust event frequency with the temporal correlation coefficient between model and station observations reaching 0.39. Simulations of dust-relevant variables such as surface wind, precipitation, and soil moisture are also evaluated. The biases including weaker surface wind speed and larger soil moisture as well as their impacts on the dust emission and transport are also noted for future model improvements. Overall, CAS-ESM2 demonstrates a good ability to simulate the spatio-temporal variations of spring dust activity over the Middle East, yet some biases remain to be improved.
... This air mass yielded the highest aerosol concentrations throughout the campaign, with particle surface area concentration modes for particles of ~0.2 µm and ~4 µm diameter, and some of the highest INP concentrations throughout the 475 campaign. This was immediately followed by an intense rainstorm later that same night (Figure 2d), which saw aerosol concentrations plummet for the following two days with the air masses coming from the northwest through (Zaitchik et al., 2007). Given this unique air mass trajectory amongst the rest of the dataset, the high INP concentrations at warmer temperatures, and 500 the "hump" in the INP spectrum that may indicate biological INP content as mentioned earlier, this data is hence discussed in further detail later in the manuscript. ...
Preprint
Full-text available
While the atmosphere in the Eastern Mediterranean is part of the dust belt, it encounters air masses from Europe, the Mediterranean Sea, and the Sahara and Arabian Deserts that bring with them a whole host of potential dust and bioaerosol compositions and concentrations via long-range transport. The consequential changes in the populations of ice-nucleating particles (INPs), aerosols that influence weather and climate by the triggering of freezing in supercooled cloud water droplets, including in the convective cloud systems in the region, are not so well understood beyond the influence of desert dust storms in increasing INP concentrations. Here, we undertook an intensive INP measurement campaign in Israel to monitor changes in concentrations and activity from four major air masses, including the potential for activity from biological INPs. Our findings show that the INP activity in the region is likely dominated by the K-feldspar mineral content, with southwesterly air masses from the Sahara Desert and easterly air masses from the Arabian Deserts markedly increasing both aerosol and INP concentrations. Most intriguingly, air masses that passed over the Fertile Crescent, a crescent-shaped region of fertile soils, agriculture, wetlands, and marshes stretching from the Nile Delta and over northern Syria to eastern Iraq, brought high INP concentrations with strong indicators of biological activity. These results suggest that the Fertile Crescent could be a sporadic source of high-temperature biological ice-nucleating activity across the region that could periodically dominate the otherwise K-feldspar-controlled INP environment and warrants further exploration in future studies in the region. This is particularly true given the ongoing desertification of the Fertile Crescent that could reveal further sources of dust and fertile soil-based INPs in the future.
... NDVI is used to monitor the green vegetation status and land cover, which comprises unstressed vegetation [23], and to evaluate ecological responses to environmental changes [40], [45], [46]. NDVI correlates with vegetation functions, such as biomass [47], [48], primary productivity [49], [50], and LAI, and can be used to investigate terrestrial vegetation and climate relationships [9], [13]- [16], [43], [51]- [54]. ...
Article
Full-text available
The spatiotemporal variability of vegetation in the Middle East was investigated for the period 2001–2019 using the Moderate Resolution Imaging Spectroradiometer (MODIS) 16-day/500 m composites of the Normalized Difference Vegetation Index (NDVI; MOD13A1). The results reveal a strong increase in NDVI coverage in the Middle East during the study period (R = 0.75, p-value = 0.05). In Egypt, the annual coverage exhibits the strongest positive trend (R = 0.99, p-value = 0.05). In Turkey, both the vegetation coverage and density increased from 2001 to 2019, which can be attributed to the construction of some of the biggest dams in the Middle East, such as the Atatürk and Ilisu dams. Significant increases in the annual coverage and maximum and average NDVI in Saudi Arabia are due to farming in the northern part of the country for which groundwater and desalinated seawater are used. The results of this study suggest that one of the main factors affecting vegetation coverage in the Middle East are governmental policies. These policies could lead to an increase in vegetation coverage in some countries such as Egypt, Saudi Arabia, Qatar, Kuwait, Iran, and Turkey.
... In recent decades, drought has turned into one of the most severe natural disasters for the ME's ecosystem in the context of global climate change (Lelieveld et al 2012, Mohammed et al 2020b. Moreover, drought has significantly affected terrestrial ecosystems, economic, social and political systems in the ME, including food security (Hameed et al 2020), hydrological processes (Bozkurt and Sen 2013), vegetation growth (Zaitchik et al 2007, Karakani et al 2021 and the extinction of many plant species (Belgacem and Louhaichi 2013). Generally, the ME's ecosystem is classified as a transitional climate pattern, located between a hot-dry and cold-humid climate pattern, which is most exposed to drought events as a result of prolonged water shortages. ...
Article
Full-text available
The primary driver of the land carbon sink is gross primary productivity (GPP), the gross absorption of carbon dioxide (CO2) by plant photosynthesis, which currently accounts for about one-quarter of anthropogenic CO2 emissions per year. This study aimed to detect the variability of carbon productivity using the Standardized Evapotranspiration Deficit Index (SEDI). Sixteen countries in the Middle East (ME) were selected to investigate drought. To this end, the yearly GPP dataset for the study area, spanning the 35 years (1982–2017) was used. Additionally, the Global Land Evaporation Amsterdam Model (GLEAM, version 3.3a), which estimates the various components of terrestrial evapotranspiration (annual actual and potential evaporation), was used for the same period. The main findings indicated that productivity in croplands and grasslands was more sensitive to the SEDI in Syria, Iraq, and Turkey by 34, 30.5, and 29.6% of cropland area respectively, and 25 31.5 and 30.5% of grass land area. A significant positive correlation against the long-term data of the SEDI was recorded. Notably, the GPP recorded a decline of >60% during the 2008 extreme drought in the north of Iraq and the northeast of Syria, which concentrated within the agrarian ecosystem and reached a total vegetation deficit with 100% negative anomalies. The reductions of the annual GPP and anomalies from 2009 to 2012 might have resulted from the decrease in the annual SEDI at the peak 2008 extreme drought event. Ultimately, this led to a long delay in restoring the ecosystem in terms of its vegetation cover. Thus, the proposed study reported that the SEDI is more capable of capturing the GPP variability and closely linked to drought than commonly used indices. Therefore, understanding the response of ecosystem productivity to drought can facilitate the simulation of ecosystem changes under climate change projections.
Article
Full-text available
In plants, stress induces changes in peroxidase enzymes, which play various physiological roles, including involvement in the development of resistance. Experiments were performed at the Elvira major Experimental Station of the NARIC Research Institute for Fruitgrowing and Ornamentals in Érd, Hungary, on two genotypes selected in Hungary (‘Alsószentiváni 117’ and ‘Milotai 10’), on the Californian-bred cultivar ‘Pedro’, and two genotypes bred in Hungary by crossing ‘Pedro’ with the two Hungarian selections: ‘Milotai intenzív’ and ‘Alsószentiváni kései’. These genotypes were chosen on the basis of frost tolerance. Measurements were made on the peroxidase activity and total polyphenol content in the leaves during sprouting (May 2nd, 12th and 20th 2016) and in the uppermost internode of the shoots (November, December 2016; January 2017). Higher peroxidase enzyme activity and polyphenol content in the uppermost internode were good indicators of the frost tolerance of the genotypes and of the stress level to which they were exposed.
Chapter
Full-text available
Due to the importance of forests in sustainable development and conservation of the global ecosystem, deforestation monitoring has become increasingly crucial in semi-arid regions. Persian Oak ( Quercus ) forests in the Zagros region, in western Iran, are one of the endangered ecosystem. The negative consequences of the destruction of forests will make life difficult in the central Iranian plateau. Development of agriculture, illegal grazing, and dust are the primary source of degradation and deforestation in the Zagros Mountains. The objective of this study was to evaluate changes in Nayangiz forest cover over two periods: (1) from 1980 to 2000, before the stricter forest law enforcement; (2) Since 2000, after enactment of strict forest laws such asinhibition of borer beetles in Oak forests and dust protection. The Landsat TM, ETM+, and OLI images were used for this purpose. After calibration, a deep learning method known as a convolutional neural network (CNN) algorithm was applied for supervised image classification. The classification accuracy was 92.2%, 94.1%, and 94.7% for all three images, respectively. Then the map of changes between each class was generated using the image classification difference method. Comparison with the changed classes shows the rates of deforestation and reforestation in each year. Before 2000, the percentage of land use change from forest cover to agricultural land was more than 37%, but in the second period this land use change (from forest cover to agricultural land) was reduced to 15%. However, a significant percentage of deforestation (23%) has been due to the effects of dust and land use change since 2000.In the study area, illegal agriculture and overgrazing led to increased dust production, which paved the way for desertification in the near future, and the mentioned factors are the factors that should be considered in order to turn forest cover into barren lands in the second period.Keywords: Land use land cover, Landsat image, convolutional neural networks, deforestation, change detection.
Article
Full-text available
Fraction of green vegetation, fg, and green leaf area index, L g , are needed as a regular space-time gridded input to evapotranspiration schemes in the two National Weather Service (NWS) numerical prediction modelsÐ regional Eta and global medium range forecast. This study explores the potential of deriving these two variables from the NOAA Advanced Very High Resolution Radiometer (AVHRR) normalized di erence vegetation index (NDVI) data. Obviously, one NDVI measurement does not allow simultaneous derivation of both vegetation variables. Simple models of a satellite pixel are used to illustrate the ambiguity resulting from a combination of the unknown horizontal (f g) and vertical (L g) densities. We argue that for NOAA AVHRR data sets based on observations with a spatial resolution of a few kilometres the most appropriate way to resolve this ambiguity is to assume that the vegetated part of a pixel is covered by dense vegetation (i.e., its leaf area index is high), and to calculate f g = (NDVI-NDVIo)/(NDVI 2 -NDVIo), where NDVIo (bare soil) and NDVI 2 (dense vegetation) are speci® ed as global constants independent of vegetation/soil type. Global (0´15ß) 2 spatial resolution monthly maps of f g were produced from a 5-year NDVI climatology and incorporated in the NWS models. As a result, the model surface ¯ uxes were improved.
Article
Full-text available
Summertime convection over Arizona typically begins in the early afternoon and continues into the night. This suggests that the evolution of the daytime planetary boundary layer is important to the development of Arizona convection. If numerical models are to provide useful guidance for forecasting convection during the monsoon, then the planetary boundary layer must be simulated as accurately as possible through utilization of the appropriate physical parameterizations. This study examines the most appropriate Pennsylvania State University-National Center for Atmospheric Research fifth-generation Mesoscale Model (MM5) planetary boundary layer parameterization(s) for deterministic and ensemble modeling of the monsoon. The four MM5 planetary boundary layer parameterizations tested are the Blackadar, Burk-Thompson, Eta, and medium-range forecast (MRF) schemes. The Blackadar and MRF planetary boundary layer schemes correctly predict the development of the deep, monsoon planetary boundary layer, and consequently do a better job of predicting the convective available potential energy and downdraft convective available potential energy, but not the convective inhibition. Because the convective inhibition is not accurately predicted, it is possible that the MM5's ability to initiate or "trigger" convection might be a limiting factor in the model's ability to produce accurate quantitative precipitation forecasts during the monsoon. Since the MM5 planetary boundary layer predicted by the Burk-Thompson and Eta schemes does not accurately reproduce the basic structure of the monsoon planetary boundary layer, their inclusion in a mixed physics ensemble is discussed.
Article
Full-text available
Modern complex societies exhibit marked resilience to interannual-to- decadal droughts, but cultural responses to multidecadal-to-multicentury droughts can only be addressed by integrating detailed archaeological and paleoclimatic records. Four case studies drawn from New and Old World civilizations document societal responses to prolonged drought, including population dislocations, urban abandonment, and state collapse. Further study of past cultural adaptations to persistent climate change may provide valuable perspective on possible responses of modern societies to future climate change.
Article
Full-text available
Substantial evolution of Normalized Difference Vegetation Index (NVDI)-derived vegetation cover (Fg) exists in the southwestern United States and Mexico. The intraseasonal and wet-/dry-year fluctuations of Fg are linked to observed precipitation in the North American monsoon system (NAMS). The manner in which the spatial and temporal variability of Fg influences the land-atmosphere energy and moisture fluxes, and associated likelihood of moist convection in the NAMS regions, is examined. For this, the regional climate model (RCM) is employed, with three different Fg boundary conditions to examine the influence of intraseasonal and wet-/dry-year vegetation variability. Results show that a strong link exists between evaporative fraction (EF), surface temperature, and relative humidity in the boundary layer (BL), which is consistent with a positive soil moisture feedback. However, contrary to expectations, higher Fg does not consistently enhance EF across the NAMS region. This is because the low soil moisture values simulated by the land surface model (LSM) yield high canopy resistance values throughout the monsoon season. As a result, the experiment with the lowest Fg yields the greatest EF and precipitation in the NAMS region, and also modulates regional atmospheric circulation that steers the track of tropical cyclones. In conclusion, the simulated influence of vegetation on land-atmosphere exchanges depends strongly on the canopy stress index parameterized in the LSM. Therefore, a reliable dataset, at appropriate scales, is needed to calibrate transpiration schemes and to assess simulated and realistic vegetation-atmosphere interactions in the NAMS region.
Article
Full-text available
Past studies have suggested that the drought of the summer of 1988 over the midwestern United States may have been caused by sea surface temperature (SST) anomalies, an evolving stationary circulation, a soil-moisture feedback on circulation and rainfall, or even by remote forcings. The relative importance of various contributing factors is investigated in this paper through the use of Goddard Earth Observing System (GEOS) GCM simulations. Seven different experiments, each containing an ensemble of four simulations, were conducted with the GCM. For each experiment, the GCM was integrated through the summers of 1987 and 1988 starting from an analyzed atmosphere in early January of each year. In the baseline case, only the SST anomalies and climatological vegetation parameters were prescribed, while everything else (such as soil moisture, snow cover, and clouds) was interactive. The precipitation differences (1988 minus 1987) show that the GCM was successful in simulating reduced precipitation in 1988, but the accompanying low-level circulation anomalies in the Midwest were not well simulated. To isolate the influence of the model’s climate drift, analyzed winds and analyzed soil moisture were prescribed globally as continuous updates (in isolation or jointly). The results show that remotely advected wind biases (emanating from potential errors in the model’s dynamics and physics) are the primary cause of circulation biases over North America. Inclusion of soil moisture helps to improve the simulation as well as to reaffirm the strong feedback between soil moisture and precipitation. In a case with both updated winds and soil moisture, the model produces more realistic evapotranspiration and precipitation differences. An additional case also used soil moisture and winds updates, but only outside North America. Its simulation is very similar to that of the case with globally updated winds and soil moisture, which suggests that North American simulation errors originate largely outside the region. Two additional cases examining the influence of vegetation were built on this case using correct and opposite-year vegetation. The model did not produce a discernible improvement in response to vegetation for the drought year. One may conclude that the soil moisture governs the outcome of the land atmosphere feedback interaction far more than the vegetation parameters. A primary inference of this study is that even though SSTs have some influence on the drought, model biases strongly influence the prediction errors. It must be emphasized that the results from this study are dependent upon the GEOS model’s identified errors and biases, and that the conclusions do not necessarily apply to results from other models.
Article
The GEWEX/ISLSCP Global Soil Wetness Project is an international initiative to enlist a spectrum of land-surface modeling centers and programs to use common global atmospheric forcing data sets to drive their respective land-surface models, in order to simulate and compare global soil wetness fields, and ultimately apply them in GCM sensitivity tests. As a participant in this project, we used the common ISLSCP atmospheric forcing data for 1987-1988 and our NCEP land-surface model (LSM) to generate global hydrological fields, including soil moisture, which cannot be obtained uniformly over large continental areas from scattered observations. We found that the simulated global hydrological cycle reasonably reflects the seasonal variation of those hydrological fields. Despite the relatively coarse 1×1 degree resolution of forcing data, the soil moisture field has a remarkable spatial variability, which is related not only to the spatial pattern of precipitation, but also to the spatial distribution of runoff and evaporation. Also, the spatial variability in vegetation and soil characteristic contributes to the variability in the surface hydrological cycle. We also compared our simulated global land-surface hydrological cycle to the 1987-1988 NCEP/NCAR reanalysis of land-surface products. It seems that, in most regions, the reanalysis of soil moisture has a larger annual cycle amplitude than the GSWP soil moisture. A more remarkable difference is that in the mid- and high-latitudes of the northern hemisphere, the reanalysis has a more uniformly distributed and wetter soil moisture than the corresponding GSWP soil moisture, which reflected the artificial climatological soil moisture damping field applied in the reanalysis system. Finally, we attempted to validate our soil moisture simulations with soil moisture measured by the Illinois Soil Moisture Network, and focused on the comparison of the area-averaged deep root zone soil moisture. It appears that the simulated soil moisture of 1987-1988 in the state of Illinois is able to reasonably capture not only the phase of the seasonal evolution, but also the amplitude of annual variation. The encouraging results indicate that a sound LSM can well simulate the natural soil moisture evolution, given accurate atmospheric forcing and reasonable specification of local vegetation and soil characteristics.
Article
Four-months-long simulations using a regional climate model covering Europe are utilized to assess the role of continental-scale vegetation cover on the seasonal evolution of soil moisture. The seasonal simulations are driven by observed lateral boundary conditions from the ECMWF (European Centre for Medium-Range Weather Forecasts) analyses, and start in April, when the soil is close to saturation. Two sensitivity experiments are conducted in which the leaf area index is doubled and halved, respectively, and these are compared to the control run.In Middle and Northern Europe an increased leaf area index leads to additional evapotranspiration and soil moisture loss from April to July. In the Mediterranean regions the situation is more complex. Initially, but only until June, the same process operates and increased vegetation leads to increased evapotransporation. In early July, however, the soil moisture content reaches a critically low value, whereupon the run with increased vegetation yields reduced evapotranspiration amounts during July.The identified reversal in sign of the vegetation / evaporation feedback is of significant conceptual and practical interest since it has the potential to shift the maximum of evapotranspiration within the summer season, and might thereby affect the seasonal cycle of precipitation.
Article
Aim Interannual land cover change plays a significant role in food security, ecosystem processes, and regional and global climate modelling. Measuring the magnitude and location and understanding the driving factors of interannual land cover change are therefore of utmost importance to improve our understanding and prediction of these impacts and to better differentiate between natural and human causes of land cover change. Despite advances in quantifying the magnitude of land cover change, the interpretation of the observed land cover change in terms of climatic, ecological and anthropogenic processes still remains a complex issue. In this paper, we map land cover change across sub‐Saharan Africa and examine the influences of rainfall fluctuations on interannual change. Location The analysis was applied to sub‐Saharan Africa. Methods Ten‐day rainfall estimates (RFE) obtained from National Oceanic and Atmospheric Administration's (NOAA) Climate Prediction Center (CPC) were used to extract information on inter and intra‐annual rainfall fluctuations. The magnitude of land cover change was quantified based on the multitemporal change vector method measuring year‐to‐year differences in bidirectional reflectance distribution function (BRDF) corrected 16‐day enhanced vegetation index (EVI) data from the Moderate Resolution Imaging Spectro‐radiometer (MODIS). Statistical models were used to estimate the relationship between short‐term rainfall variability and the magnitude of land cover change. The analysis was stratified first by physiognomic vegetation type and second by chorological data on species distribution to gain insights into spatial variations in response to short‐term rainfall fluctuations. Results The magnitude of land cover change was significantly related to rainfall variability at the 5% level. Stratification considerably strengthened the relationship between the magnitude of change and rainfall variability. Explanatory power of the models ranged from R ² = 0.22 for the unstratified model to 0.40–0.96 for the individual models stratified by patterns of species distribution. The total variability explained by the combined models including the influence of rainfall and differences in vegetation response ranged from 22% for the model not stratified by vegetation to 76% when stratified by chorological data. Main conclusions Using this methodology, we were able to measure the contribution of natural variation in precipitation to land cover change. Several ecosystems across sub‐Saharan Africa are highly sensitive to short‐term rainfall variability.
Article
The definition of desertification accepted in the ad hoc conference held by UNEP in Nairobi in 1977 and confirmed at the Earth Summit on Environment and Development held in Rio de Janeiro in 1992 is: ‘arid, semi-arid and dry-subhumid land degradation’. There is no global long-term trend in any rainfall change over the period of instrumental record (c. 150 years), but there has been an increase of 0·5°C in global temperature over the past 100 years. This increase seems partly due to urbanization, as there is no evidence of it resulting from atmospheric pollution by CO2and other warming gases (SO2, NO2, CH4, CFH etc.). On the other hand, the thermal increase is uneven, increasing with latitudes above 40° N and S. The increase is only slight or non-existent in subtropical and inter-tropical latitudes where most arid and semi-arid lands lie. This, incidently, is consistent with Global Circulation Models (GCM) — derived scenarios. The study of tree-rings, lake level fluctuations and pollen analysis confirm the existence of climatic fluctuations, but with no long-term trends over the past 2000 years.