SlideShare a Scribd company logo
1 of 10
Download to read offline
Analysis of teleconnections between AVHRR-based sea surface temperature and
vegetation productivity in the semi-arid Sahel
Silvia Huber ⁎, Rasmus Fensholt
Dept. of Geography and Geology, University of Copenhagen, Øster Voldgade 10, DK-1350 Copenhagen, Denmark
a b s t r a c ta r t i c l e i n f o
Article history:
Received 15 February 2011
Received in revised form 11 July 2011
Accepted 13 July 2011
Available online 13 September 2011
Keywords:
Climate indices
NDVI
Oceanic forcing
Remote sensing
Sahel
SST
Vegetation dynamics
Vegetation productivity across the Sahel is known to be affected by a variety of global sea surface temperature
(SST) patterns. Often climate indices are used to relate Sahelian vegetation variability to large-scale ocean–
atmosphere phenomena. However, previous research findings reporting on the Sahelian vegetation response
to climate indices have been inconsistent and contradictory, which could partly be caused by the variations in
spatial extent/definitions of climate indices and size of the region studied. The aim of this study was to analyze
the linkage between climate indices, pixel-wise spatio-temporal patterns of global sea surface temperature
and the Sahelian vegetation dynamics for 1982–2007. We stratified the Sahel into five subregions to account
for the longitudinal variability in rainfall. We found significant correlations between climate indices and the
Normalized Difference Vegetation Index (NDVI) in the Sahel, however with different magnitudes in terms of
strength for the western, central and eastern Sahel. Also the correlations based on NDVI and global SST
anomalies revealed the same East–West gradient, with a stronger association for the western than the eastern
Sahel. Warmer than average SSTs throughout the Mediterranean basin seem to be associated with enhanced
greenness over the central Sahel whereas colder than average SSTs in the Pacific and warmer than average
SSTs in the eastern Atlantic were related to increased greenness in the most western Sahel. Accordingly, we
achieved high correlations for SSTs of oceanic basins which are geographically associated to the climate
indices yet by far not always these patterns were coherent. The detected SST–NDVI patterns could provide the
basis to develop new means for improved forecasts in particular of the western Sahelian vegetation
productivity.
© 2011 Elsevier Inc. All rights reserved.
1. Introduction
The development and prosperity of the African Sahel, a semi-arid
transition zone between the Sahara desert to the north and the humid
tropical savanna to the south, largely depend on the rainfall regime.
Vegetation growth is limited to the short rainy season with July to
September being the months of highest rainfall (Nicholson et al.,
2000). The vegetation phenology thereby closely responds to the
seasonal cycle of rain with vegetation productivity taking place
mainly during the months with rainfall (Herrmann et al., 2005;
Hickler et al., 2005; Huber et al., 2011; Prince et al., 1998). However,
precipitation in the Sahel shows typically large interannual and
decadal variability. This variability was impressively manifested
starting in the mid 1960s, when rainfall amounts continuously
declined across the Sahel culminating in a widespread drought in
1982–1985. But since then, rainfall has steadily increased together
with the vegetation resource base. This greening up has been
observed from satellite based Normalized Difference Vegetation
Index (NDVI) time series (Anyamba and Tucker, 2005; Eklundh and
Olsson, 2003; Herrmann et al., 2005; Heumann et al., 2007; Seaquist et al.,
2008). Due to the close coupling between rainfall and vegetation growth
in the Sahel, rainfall variability explains a significant part of this increase
in NDVI (Fensholt and Rasmussen, 2011; Herrmann et al., 2005; Hickler
et al., 2005;Huber et al., 2011;Nicholsonetal., 1990).NDVI has,despite of
the index simplicity, shown to be a reliable proxy for vegetation vigor and
productivity in the Sahel (Prince, 1991a,b; Rasmussen, 1992; Tuckeret al.,
1985).
The reasons for the large interannual and decadal fluctuations in
rainfall are still not entirely clear but the early works of Folland et al.
(1986), Lamb (1978) and Palmer (1986) highlighted that rainfall
variability across the Sahel is related (teleconnected) with regional and
global sea surface temperature (SST) conditions. These include
interannual and decadal temperature anomalies in the Atlantic (e.g.,
Biasutti et al., 2008; Palmer, 1986; Shanahan et al., 2009; Ward, 1998),
the Pacific (e.g., Caminade and Terray,2010; Janicot et al., 1998; Mohino
et al., 2010), the Indian Ocean (e.g., Bader and Latif, 2003; Giannini et al.,
2003; Lu, 2009) and the Mediterranean (e.g., Philippon et al., 2007;
Raicich et al., 2003; Rowell, 2003). Often climate indices are used to
relate Sahelian rainfall or vegetation variability to large-scale ocean–
atmosphere phenomena. Climate indices reflect the essential elements
Remote Sensing of Environment 115 (2011) 3276–3285
⁎ Corresponding author. Tel.: +45 35322 499.
E-mail addresses: huber.silvia@gmail.com (S. Huber), rf@geo.ku.dk (R. Fensholt).
0034-4257/$ – see front matter © 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.rse.2011.07.011
Contents lists available at SciVerse ScienceDirect
Remote Sensing of Environment
journal homepage: www.elsevier.com/locate/rse
of climate and its fluctuations through time. They can either be very
simple, for example only calculated from SSTs, or integrate a larger
number of climate variables, such as temperature, pressure or surface
winds. Some relate to interannual climate patterns, while others reflect
decadal to inter-decadal phenomena. The El Niño-Southern Oscillation
(ENSO) is one of the best known indices of interannual climate
variability. However, its impact on the Sahelian vegetation is still
debated. While several studies have reported no or only a minor
influence (Anyamba et al., 2001; Anyamba and Eastman, 1996;
Anyamba and Tucker, 2005; Philippon et al., 2007; Propastin et al.,
2010), others have shown that the Sahelian climate is related to ENSO
events (Camberlin et al., 2001; Oba et al., 2001; Ward, 1998). Other
climate indices that were associated to the Sahelian vegetation
dynamics are the North Atlantic Oscillation (NAO), the Pacific Decadal
Oscillation (PDO) and the Indian Ocean Dipole (IOD). Oba et al. (2001)
attributed in their study large parts of the interannual variation of
vegetation productivity during the 1980s to the NAO while Wang
(2003) could not find a consistent relationship. The influence of the
PDO and the IOD on the Sahelian climate is less studied. Brown et al.
(2010) found significant relationships between the PDO and the
growing season in West Africa while the IOD was found to have
virtually no influence in the Sahel. Also Williams and Hanan (2011)
found only weak responses across the Sahelian zone in their study
when looking at the IOD and the Multivariate ENSO Index (MEI)
individually but not when investigating interacting effects of the two
indices. All these findings suggest some contradiction in the Sahelian
response to changes as characterized by climate indices, which could
partly be caused by the variations in spatial extent respectively
definitions of climate indices and size of the region studied.
The goal of this paper is to analyze the linkage between these
climate indices, pixel-wise spatio-temporal patterns of global sea
surface temperature and the Sahelian vegetation dynamics. In this
analysis the Sahel, an ecoregion covering an area of more than
3 million km2
, was divided into five subregions because it has been
shown that the Atlantic section of the Sahel is only weakly related
with the rest of the zone (Moron, 1994). Moreover the definition of
the “Sahel window” varies considerably among different studies, in
particular in the longitudinal extent. For example Giannini et al.
(2003) referred to the 20°W–35°E/10–20°N extent, Mohino et al.
(2010) used the area 15°W–15°E/10–17°N, Jarlan et al. (2005) studied
the window 18°W–18°E/12–18°N and Rowell (2003) defined the area
16.875°W–35.625°E/11.25–18.75°N as the Sahel. Here we used the
longitudinal extent defined in the study of Giannini et al. (2003)
(20°W–35°E) but stratified the Sahel into five subregions as displayed
in Fig. 1 to account for the longitudinal variability in rainfall.
Using Pearson's correlation analysis between climate indices,
remotely sensed time series of SST and NDVI corrected for seasonality
and trend, the following research questions were investigated: a) is
there a statistically significant relationship between global SST
anomalies/climate indices and NDVI in the Sahel and are the spatial
patterns consistent; b) which ocean regions explain most of the NDVI
dynamics across the subdivided Sahel; c) which time lags are involved
in the interrelations between SST/climate indices and NDVI and d) do
certain areas in the Sahel exhibit stronger teleconnections to SST
anomalies than others. The spatio-temporal correlation analysis also
including sea surface temperature may help to unveil new options to
develop tools for forecasting the Sahelian resource base, which is
largely dependent on rainfall.
2. Data
Four climate indices and two different satellite-based gridded
monthly time series (NDVI and SST) were used in this study. High-
temporal continuous EO-based NDVI and SST is currently available from
1982 to 2007 (full year coverage=312 months). This time period thus
determines the temporal extent of the analyses performed.
2.1. Climate indices
We used four climate indices based on global climate variations
from various oceanic regions (Fig. 2).
2.1.1. Indian Ocean Dipole (IOD)
The Indian Ocean Dipole (IOD) is an interannual climate
phenomenon in the tropical Indian Ocean, first defined by Saji et al.
(1999). In the positive phase of the IOD, trade winds are stronger than
usual and cooler-than-average sea-surface temperatures are preva-
lent across the eastern tropical Indian Ocean, near Indonesia and
Australia. To the west, near Madagascar, waters are warmer than
average and convection is intensified. These patterns are reversed
during the IOD's negative phase (UCAR, 2011). Intensity of the IOD is
represented by anomalous SST gradients between the western
equatorial Indian Ocean (50E–70E and 10S–10N) and the south
eastern equatorial Indian Ocean (90E–110E and 10S–0N). We used
the IOD based on HadlSST SST (monthly from 1982 to 2007)
downloaded from: http://www.jamstec.go.jp/frsgc/research/d1/iod/.
2.1.2. Multivariate ENSO Index (MEI)
The Multivariate El Niño-Southern Oscillation (ENSO) Index (MEI)
is a measure to monitor the strength of ENSO conditions. ENSO is the
most important coupled ocean–atmosphere phenomenon to cause
global climate variability on interannual time scales. MEI is derived by
combining the six main observed variables in the tropical Pacific: SST,
sea-level pressure, surface winds, surface air temperature, cloudiness
and precipitation (Wolter and Timlin, 1998). The MEI key regions for
the variable measurements can be seen in Wolter and Timlin (1998,
Fig. 1). MEI is favored over conventional indices, since it combines the
significant features of all observed surface parameters in the tropical
Pacific (Wolter and Timlin, 1998). MEI was obtained from: http://
www.esrl.noaa.gov/psd/data/climateindices/.
2.1.3. North Atlantic Oscillation Index (NAO)
The North Atlantic Oscillation (NAO) is typically measured
through variations in the normal pattern of lower atmospheric
Fig. 1. Map of the average annual rainfall in the Sahel (1996–2007) and the areas of study: two regions have been selected in the western (W1, W2) and central Sahel (C1, C2) and
one region in the eastern Sahel (E1).
3277S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
pressure over Iceland and higher pressure near the Azores and Iberian
Peninsula (UCAR, 2011). A positive NAO index refers to stronger than
usual subtropical high pressure center around the Azores and a deeper
than normal Icelandic low, with increased pressure generating more
and stronger winter storms crossing the Atlantic Ocean on a more
northerly track. Europe tends towards warm and wet winters while
northern Canada and Greenland will usually have cold and dry
winters, with the eastern United States generally experiencing mild,
wet winter conditions.
A negative NAO index (when there is less difference than usual in
pressure across the two regions) features a weakened Atlantic storm
track, a greater risk for Arctic outbreaks of cold air across the
northeastern United States and northern Europe, and moist air to the
Mediterranean (UCAR, 2011). NAO varies from year to year but has a
roughly decadal pattern with a dominant period of 12 years (Deser
and Blackmon, 1993). The index can be obtained from: http://www.
esrl.noaa.gov/psd/data/climateindices/.
2.1.4. Pacific Decadal Oscillation (PDO)
The Pacific Decadal Oscillation (PDO) is a multi-decadal pattern
of climate variability across the North Pacific Ocean (Mantua et al.,
1997). The positive (warm) mode of PDO features colder than
average SSTs in the central North Pacific along a narrow band of
warmer SSTs along the west coast of North America and in the
eastern tropical Pacific. During the negative (cool) phase of PDO,
the opposite has been observed: a warm pool of sea surface waters
in the central north Pacific and cold SSTs along the west coast
(Mantua et al., 1997).
Each phase typically persists for 20 to 30 years, with a warm phase
predominating since the late 1970s. The PDO may be related to ENSO,
but differs mainly because the timescale for the PDO is much longer
(several decades) and because the PDO more clearly involves the
extratropical Pacific and the Aleutian Low pressure system (UCAR,
2011). Even though PDO is mirroring decadal patterns, it involves
sufficient interannual variability to relate to productivity in Africa
(Brown et al., 2010). The index is available at http://jisao.washington.
edu/pdo/PDO.latest.
2.2. Normalized Difference Vegetation Index (NDVI)
We used AVHRR (Advanced Very High Resolution Radiometer)
GIMMS (Global Inventory Modeling and Mapping Studies) NDVI data
as a proxy for vegetation productivity. Monthly maximum NDVI
composites with an 8 km spatial resolution used in this study were
processed by the GIMMS Group at NASA's Goddard Space Flight
Center, as described by Tucker et al. (2005) and Pinzon et al. (2005).
More details about the binning and band calibration can be found in
James and Kalluri (1994), Vermote and Kaufman (1995) and Los
(1998). No atmospheric correction is applied to the GIMMS data
except for volcanic stratospheric aerosol periods (1982–1984 and
1991–1994) (Tucker et al., 2005). A satellite orbital drift correction is
performed using an empirical mode decomposition (EMD) transfor-
mation method of Pinzon et al. (2005) removing common trends
between time series of solar zenith angle (SZA) and NDVI. The original
16-bit GIMMS NDVI was converted into real NDVI values (range −1 to
1) for further analysis.
2.3. Sea surface temperature (SST)
NOAA Optimum Interpolation (OI) SST v2 data, provided by the
NOAA-CIRES Climate Diagnostics Center, Boulder, USA were used in
this paper (http://www.esrl.noaa.gov/psd/data/gridded/data.noaa.
oisst.v2.html). The NOAA SST OIv2 product integrates both in situ
and satellite data from November 1981 to the present at 1.0° spatial
resolution globally in degrees Celsius (Reynolds et al., 2002). The in
situ SST data are determined from observations from ships and buoys
(Reynolds et al., 2002). Satellite data is also obtained from the AVHRR
instrument. The NOAA OI.v2 SST monthly fields are derived by a linear
interpolation of the weekly optimum interpolation (OI) version 2
fields to daily fields and then averaging the daily values over a month.
More details about the product can be found in Reynolds et al. (2002).
3. Methods
The boundary of the Sahel used in this paper (indicated on the
maps in Figs. 1, 4 and 5) is defined by Le Houérou (1989) using the
IOD
-3.5
-2.5
-1.5
-0.5
0.5
1.5
2.5
3.5
1982 1985 1988 1991 1994 1997 2000 2003 2006
Year
1982 1985 1988 1991 1994 1997 2000 2003 2006
Year
1982 1985 1988 1991 1994 1997 2000 2003 2006
Year
1982 1985 1988 1991 1994 1997 2000 2003 2006
Year
NormalizedIndex
-3.5
-2.5
-1.5
-0.5
0.5
1.5
2.5
3.5
NormalizedIndex
-3.5
-2.5
-1.5
-0.5
0.5
1.5
2.5
3.5
NormalizedIndex
-3.5
-2.5
-1.5
-0.5
0.5
1.5
2.5
3.5
NormalizedIndex
MEI
NAO PDO
Fig. 2. Monthly climate indices Indian Ocean Dipole (IOD), Multivariate ENSO Index (MEI), North Atlantic Oscillation Index (NAO) and Pacific Decadal Oscillation (PDO) from 1982 to
2007.
3278 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
150 (to the north) and 700 mm (to the south) annual rainfall isohyets
as borders of the Sahelian area based on rainfall mean annual values of
the Rainfall Estimate (RFE) blended gage-satellite rainfall product
(NOAA Climate Prediction Center) for 1996–2007 (Herman et al.,
1997). Within this boundary we defined five subregions (Fig. 1): two
in the Western (W1, W2), two in the Central (C1, C2) and one in the
Eastern Sahel (E1) to account for different rainfall regimes in the East–
West orientation at the interannual scale. The longitudinal division
was partly based on the analysis conducted by Moron (1994) and
Lebel and Ali (2009). Three subregions (W1, W2 and C2) are 10° wide
in longitude and C1 and E1 were chosen to be 15° wide in latitude. It
was decided to define C1 with a larger width to keep Chad incl. the
Marra mountains as well as the Sahelian part of Sudan as an entity. E1
was enlarged to 15° since rainfall anomalies of the most eastern Sahel
(35–40°E) and the region between 25° and 35°E are highly correlated.
3.1. Preprocessing
All analyses were performed on linearly detrended NDVI and SST
time series. Further, for both time series standardized anomalies were
computed to remove the seasonal component from the data. The
original 15-day NDVI composite data were aggregated to months
using a maximum value composite approach to further reduce the
influence from clouds (Holben, 1986). Monthly anomalies for each
26-year data record (1982–2007=312 images) were obtained by
computing the median value for each pixel for each month
(=climatology value) which was then subtracted from each image.
Median values rather than mean values were used because the time
series of this study are shorter than the standard 30-year baseline
period defined by the World Meteorological Organization (WMO) to
calculate climatology values. The SST time series as well as the climate
indices were additionally smoothed using a moving average filter over
a 3-month period (Jan–March, Feb–April etc.) to reduce noise
(Plisnier et al., 2000).
From the NDVI time series mean values for the period July–
September (termed JAS hereafter) were calculated for each of the five
subregions in the Sahel. Finally, area-average mean JAS NDVI were
extracted for all subregions, termed hereafter NDVI Anomaly Indices
(NAI) (Fig. 3).
3.2. Teleconnections
Pixel-wise Pearson correlation coefficients were calculated be-
tween the four climate indices (IOD, MEI, NAO and PDO), global SST
anomalies (3-month means) and NAI values for each of the five
subregions from 1982 to 2007. Since changes in ocean-atmospheric
patterns can affect the Sahelian rainfall regime and hence the
vegetation development with a delay of several months, lagged
correlation was used, with climate indices and SST observed at steps
0 to 9 months prior to the NDVI observations in the Sahel. Next,
oceanic regions (all substantially larger than the Sahel) with highest
positive or negative correlations with JAS NDVI were identified and
three indices of SST anomalies extracted (Fig. 4). Two indices were
extracted in the northern Atlantic Ocean. The first one for January–
March mean SSTs as displayed in Fig. 6b: JFM) (N3.5 million km2
) and
the second one for June–August mean SSTs as illustrated in Fig. 6a:
JJA) (N4.5 million km2
). The third index covers an area of more than
5 million km2
in the equatorial Pacific and was extracted from March–
May mean SSTs (Fig. 6a: MAM). These three SST indices (JFM Atlantic,
JJA Atlantic and MAM pacific) were correlated with JAS NDVI across
the Sahel and the results compared with the commonly used climate
indices. To explore combined effects of the indices, partial correlation
analysis was applied.
4. Results and discussion
Correlations of JAS NDVI anomalies (NAI) between the different
sub-regions of the Sahel confirm that the NDVI dynamics of the
Atlantic section (W1) is to a certain degree decoupled from the rest of
the region (Table 1) as also suggested by (Moron, 1994). The
subregion W2 is closer related to the central Sahel (C1) than to W1.
The highest correlation was found between W2 and C1 (r=0.78).
4.1. Correlations between climate indices and Sahelian NDVI dynamics
The analysis revealed that the western and central Sahel is better
correlated to predefined climate indices (IOD, MEI, NAO and PDO)
than the eastern Sahel. Most significant correlations were achieved for
MEI and PDO (Fig. 5a and b). In particular for Senegal, southern
Mauritania and western Mali we found highly significant r-values for
MEI with the highest number of significant pixels for the periods AMJ
and MJJ. Brown et al. (2010) also reported significant correlations
between MEI aggregated over June to August (JJA) and the start of the
growing season (SOS) for the westernmost Sahel, however in their
study they report positive correlations while we found values of the
opposite sign. Other studies report only modest to small ENSO
response of NDVI for the Sahel (Anyamba and Eastman, 1996;
Philippon et al., 2007; Williams and Hanan, 2011).
Also for PDO significant negative correlations were found
throughout the western to central Sahel; in particular areas of Mali
Fig. 3. Area-averaged JAS NDVI standardized anomalies (NAI) obtained from linearly detrended time-series from 1982 to 2007 for the five Sahelian subregions as outlined in Fig. 1.
3279S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
and Burkina Faso were related to the index (Fig. 5b). The highest
anticorrelations for PDO and JAS NAI were found when averaging PDO
over JFM and JAS. In contrast to the results obtained for MEI and PDO,
we found some positive correlations when testing associations
between JAS NAI and IOD and NAO indices, respectively, (Fig. 5c
and d); yet a smaller number of pixels with significant r-values was
found as compared to MEI and PDO which is in accordance with the
findings of Brown et al. (2010).
4.2. Correlations between global SST and Sahelian NDVI dynamics
The pixel-wise correlation analysis revealed several significant
relationships between SST and vegetation dynamics in the Sahel
(Fig. 6). In general larger oceanic areas with significant correlations
were found for the western and central Sahel than for the eastern
Sahel. This is in accordance with the findings presented in the
previous section. Subsequently we report teleconnections between
JAS NAI and SST anomalies for the five different subregions. Only large
scale patterns of significant correlations that are persistent over at
least three aggregated time periods will be commented on.
4.2.1. Sahel subregion W1 (20–10° W)
The results obtained for W1 differ from the other four subregions,
corroborating the fact that the Atlantic section of the Sahel is less well
related to the rest of the ecoregion as described above. Compared to
the other subregions, W1 in general shows stronger associations to
oceanic temperature anomalies reflected by more pixels with
significant r-values (Fig. 6a).
In particular large areas of the Pacific Ocean emerge with significant
correlations (p≤0.01), both positive and negative. Anticorrelations are
very distinct in the tropical Pacific, with decreasing r-values for shorter
lag-periods while the areas to the north and south of the tropical Pacific
exhibit positive correlations. For the Philippine Sea also positive
relationships were found between SST and JAS NAI anomalies for the
periods JFM, FMA and MAM. The anticorrelations imply that lower than
average SSTs in the tropical Pacific Ocean are statistically related to
higher boreal summer precipitation in the West Sahel (W1). These
results are consistent with other research. Ward (1998) found similar
patterns for “no dipole years”, i.e., years with rainfall anomalies of the
same sign in the Sahel and the Guinean Coast at high-frequency
timescale, when correlating JAS SST and Sahelian rainfall. Higher
photosynthetic activity over the Sahel was also associated to negative
summer SST anomalies in the tropical eastern Pacific in the study of
Philippon et al. (2007). However, in this study we could link winter/
spring SSTs to summer NDVI in the Sahel. This time lag could be
important for predictive purposes. The strong pattern we obtained in
the Pacific Ocean would suggest that there is also a link between PDO
and JAS NAI in the Western Sahel, but this was not the case (Fig. 5b).
For shorter lag-periods (starting with MJJ) high anticorrelations
emerged in the Indian and Southern Ocean, with r-values between −
0.60 and −0.68. Jury and Mpeta (2009) showed an association
between the westerly wind anomalies and sea-level pressure of the
South Indian Ocean and Sahel rainfall with a 12-month lead time. The
Indian Ocean in general plays an important role in forcing the Sahelian
climate, in particular at decadal time scales, as reported in various
studies (e.g., Bader and Latif, 2003; Giannini et al., 2003; Hagos and
Cook, 2008). Philippon et al. (2007) however did not find any
significant signal over the tropical Indian Ocean. They argue in their
study that this is merely attributed to the time period that is restricted
to the recent two decades, which does not capture longterm decadal
variability. Even though the time scale covered in this study does not
allow detecting decadal or multi-decadal patterns, patterns in the
Indian Ocean showed to be associated with JAS NAI.
Together with the occurrence of significant patterns in the Indian
Ocean, positive correlations occur in the western part of the North
Atlantic Ocean, from the MJJ period onwards (p≤0.01, with highest
correlations for JAS (r=0.84)) characterizing a large region of pixels
(horseshoe shaped) stretching from just off the coast of Senegal to the
Atlantic ocean west of the European continent. This spatial extent
corresponds well with the cold ocean current associated with one of
the five major existing gyres (the North Atlantic gyre) (Heinemann
and Open University Course Team, 1989).
However, even though we found these positive correlations in the
northern Atlantic and a weak correlation (r=0.35) between JAS NAO
and the JJA Atlantic index (extracted in this study and described in
Section 3.2), almost no significant relationships were found when
correlating NAO with JAS NAI for the subregion W1.
4.2.2. Sahel subregion W2 (10–0° W)
For JAS NAI of W2 in particular two oceanic areas emerged with
highly significant negative correlations (p≤0.01): an area of the
Atlantic Ocean for longer lag-periods (JFM, FMA and MAM) and an
area of the Indian Ocean with a maximum r value of −0.8 for SST
anomalies averaged for MJJ (Fig. 6b). The strong link between
northern Atlantic SSTs and NDVI anomalies in the subregion W2 is
in agreement with the findings of the correlation analysis between JAS
NAI and NAO. Also for NAO we found significant (but weak)
correlations with W2 NDVI for long lag-periods, in particular JFM,
Fig. 4. Extracted SST anomaly indices for the Atlantic averaged over January–March (JFM) and June–August (JJA), respectively, and the equatorial Pacific for March–May (MAM).
Table 1
Pearson's correlation coefficients between area-averaged JAS NDVI anomaly indices
(NAI) for the five subregions in the Sahel.
Subregion W1 W2 C1 C2 E1
W1 [20–10°W] 0.44 0.17 0.12 0.03
W2 [10–0°W] 0.44 0.78 0.43 0.26
C1 [0–15°E] 0.17 0.78 0.63 0.35
C2 [15–25°E] 0.12 0.43 0.63 0.63
E1 [25–40°E] 0.03 0.26 0.35 0.63
3280 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
even though NAO represents pressures which are not directly
comparable to SST patterns.
Compared to the results of W1 a general decrease in the
geographical extent of significant pixels can be observed. In particular
the association to the Pacific Ocean is much weaker. For example, W2
is the only subregion for which almost no significant relationships
were detected in the Western Pacific. But still we identified significant
negative correlations between JAS NAI of this region (and also C1 and
C2, but not for W1) and the PDO. Interestingly some SST patterns in
the North Pacific and the Aleutian Islands temporally coincide with
the strongest NAI–PDO relationships. PDO has been shown to clearly
involve the extratropical Pacific and the Aleutian low pressure system
(UCAR, 2011).
Finally, while we detected for W1 strong positive correlations from
MJJ onwards in the northern Atlantic, for W2 the analysis revealed
negative correlations for JFM until AMJ. The reason for this pattern
needs to be further investigated.
4.2.3. Sahel subregion C1 (0–15° E)
Similar to W2, also for the JAS NDVI dynamics of C1 anticorrelations
were discovered in the Indian Basin with max. r-values of −0.63 for the
period MAM (Fig. 6c). Starting with AMJ the Eastern Mediterranean Sea
showed increasing associations with NAI. The findings that the
Mediterranean plays an important role in the Sahelian climate are
consistentwiththoseof other studies (Raicich et al., 2003; Rowell, 2003)
and suggest that a warming of the Mediterranean is often associated
with enhanced Sahelian rainfall and hence increased vegetation growth.
The main reason for the regional teleconnection leading to additional
moisture over the Sahel in warm Mediterranean years is increased
evaporation over the sea surface leading to an enhanced moisture
content of the air that is advected southwards across the eastern Sahara
into the Sahel (Rowell, 2003). During winter, the whole Mediterranean
is influenced by westerlies (Raicich et al., 2003). During summer the
westerlies are weaker and a meridional regime develops, especially over
the eastern Mediterranean basin (=Etesian wind regime) which
connects the eastern Mediterranean area with the sub-Saharan
ecoregion (Raicich et al., 2003). For the period JFM and FMA the
western Pacific also partly contributes to JAS NDVI variability.
4.2.4. Sahel subregion C2 (15–25° E)
For C2 positive relationships were found between JAS NAI and SST
anomalies in the South West pacific for longer lag-periods (JFM, FMA,
MAM) (Fig. 6d). Only in July–September (JAS) the SST anomalies
measured in the SE Mediterranean and the Red Sea corresponded well
with NAI (r=0.65). As for C1, the study also revealed significant
positive correlations for the western Pacific for JFM, FMA and MAM.
4.2.5. Sahel subregion E1 (25–40° E)
With the JS NAI of this region the SST anomalies of only a few
oceanic areas were significantly correlated (Fig. 6e). Teleconnections
were found almost exclusively in the Atlantic Ocean and Western
Pacific for longer lag-periods. The last two periods analyzed (JJA and
Fig. 5. Maps of significant r-values from correlation analyses between JAS NDVI anomalies and a) the Multivariate ENSO Index (MEI) averaged over May–July, b) the Pacific Decadal
Oscillation (PDO) averaged over July–September, c) the North Atlantic Oscillation (NAO) averaged over January–March and d) the Indian Ocean Dipole (IOD) averaged over April–
June from 1982 to 2007.
3281S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
Fig. 6. Maps of correlation coefficients (r) between the Sahel NDVI anomaly index (NAI) extracted from the five subregions: a) W1, b) W2, c) C1, d) C2 and e) E1) and mean SST
anomalies from 1982 to 2007 for different time lags (e.g., JFM: January–March). Only correlations between the 1 and 5% confidence levels are shown.
3282 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
JAS) revealed positive associations between the SST of the Red Sea and
the NDVI dynamics of the Eastern Sahel as it was the case for the
neighboring subregion C2. These findings correspond with the climate
indices analysis, which revealed hardly any significant correlations for
the most eastern Sahel.
4.3. Joint correlations between climate/SST indices and Sahelian NDVI
dynamics
The results obtained from a partial correlation analysis are
presented in Fig. 7. It can be seen that in certain hotspot regions
across the Sahel up to 50% of the interannual variability in NDVI can be
explained by the combined indices. With the combined JFM Atlantic
and MAM Pacific indices hotspots of high r2
-values were mapped
across the Sahel, except for C2 (Fig. 7b). Interestingly, highest
correlations were achieved for C2 with MJJ MEI and JAS PDO
(Fig. 7a). When the JJA Atlantic index was used in combination with
the MAM Pacific index, in particular large parts of the NDVI variability
of the western Sahel can be explained (Fig. 7c). In Senegal, Mali and
Mauritania large areas are mapped with r2
-values between 0.4 and
0.5. These results illustrate that a larger amount of the JAS NDVI
variability in the western Sahel can be explained when using
extracted SST indices as compared to the well known MEI and PDO.
Yet the latter were able to explain more of the JAS NDVI variability in
subregion C2.
5. Conclusions
Using EO-based time series of SST and NDVI covering 1982–2007
provides the possibility of performing analyses based on observations
continuous in space and time but on the other hand the time record of
the AVHRR sensors is not long enough to discover decadal patterns.
We therefore focused on interannual patterns in this study. From this
analysis we conclude that significant correlations exist between
global SST anomalies and Sahelian NDVI, however with different
characteristics for western, central and eastern Sahel. Whereas the
vegetation productivity in the western Sahel could be associated with
large oceanic areas of the Pacific, the Atlantic as well as the Indian
Ocean, for the eastern Sahel only small areas in the Atlantic were
found to be significantly related to dynamics in NDVI. The Eastern
Mediterranean emerged only with significant r-values when related
to NDVI in the central Sahel. Warmer than average SSTs throughout
the Mediterranean basin seem to be associated with enhanced
greenness over the central Sahel whereas colder than average SSTs
in the Pacific and warmer than average SSTs in the eastern Atlantic
were related to increased greenness in the most western Sahel.
The correlations based on NAI and climate indices revealed the
same East–West gradient, with stronger associations for the western
than the eastern Sahel. Accordingly, we achieved high correlations for
SSTs of oceanic basins which are associated to the indices (e.g., for
W1: MEI and equatorial Pacific) yet by far not always. For instance
relating IOD with Sahelian NDVI did not result in higher correlations
even though the SSTs of the Indian Ocean played an important role for
the NDVI dynamics in W1, W2 and C1. This result may be explained by
the fact that IOD is defined as a gradient in the equatorial Indian
Ocean. However, for W2 and C1 the most significant correlations were
found in the southern Indian Ocean, away from the Equator and for
W1, the correlations are significant over the Equator but show the
same signed structure and no gradient.
For NDVI of W1 the study showed a strong association to the SST
anomalies of the Pacific Ocean but when relating PDO to JAS NAI this
link could not be reproduced for W1.
Overall, these large scale climate indices do have predictive power
but they might be defined too broad thereby suppressing predictive
capabilities of more localized areas like it seems to be the case for the
Atlantic SST anomaly we found along the Senegal–Europe area for JAS.
This is reflected in the findings of the correlation analysis based on
combined indices. It illustrates that with extracted SST indices for
Fig. 7. Maps of joint correlations (r2
) from partial correlation analysis of JAS NDVI anomalies and a) the Multivariate ENSO Index (MEI) averaged over May–June and the Pacific
Decadal Oscillation (PDO) averaged over June–August, b) the SST indices extracted from the Atlantic and Pacific for January–March and March–May, respectively, and c) the SST
indices extracted from the Atlantic and Pacific for June–August and March–May, respectively.
3283S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
specific oceanic areas the percentage of explained NDVI variability can
be increased and extended to larger areas as compared to traditional
climate indices. However, in this paper we only investigated linear
SST–NDVI relationships. Also non-linear features or interferences
between climate patterns might play a role and this should be
considered in future studies.
The detected SST–NDVI patterns could provide the basis to
develop new means for improved forecasts in particular of the
western Sahelian vegetation resource base for pastoralism and
agricultural production.
Acknowledgements
The authors thank the NASA Global Inventory Modeling and
Mapping Studies (GIMMS) group for producing and sharing the
AVHRR GIMMS NDVI data set as well as the NOAA-CIRES Climate
Diagnostics Center, Boulder, USA, for providing the NOAA Optimum
Interpolation (OI) SST v2 data. Thanks to two anonymous reviewers
for their helpful comments.
References
Anyamba, A., & Eastman, J. R. (1996). Interannual variability of NDVI over Africa and its
relation to El Nino Southern Oscillation. International Journal of Remote Sensing, 17,
2533–2548.
Anyamba, A., & Tucker, C. J. (2005). Analysis of Sahelian vegetation dynamics using
NOAA-AVHRR NDVI data from 1981–2003. Journal of Arid Environments, 63,
596–614.
Anyamba, A., Tucker, C. J., & Eastman, J. R. (2001). NDVI anomaly patterns over Africa
during the 1997/98 ENSO warm event. International Journal of Remote Sensing, 22,
1847–1859.
Bader, J., & Latif, M. (2003). The impact of decadal-scale Indian Ocean sea surface
temperature anomalies on Sahelian rainfall and the North Atlantic Oscillation.
Geophysical Research Letters, 30.
Biasutti, M., Held, I. M., Sobel, A. H., & Giannini, A. (2008). SST forcings and Sahel rainfall
variability in simulations of the twentieth and twenty-first centuries. Journal of
Climate, 21, 3471–3486.
Brown, M. E., de Beurs, K., & Vrieling, A. (2010). The response of African land surface
phenology to large scale climate oscillations. Remote Sensing of Environment, 114,
2286–2296.
Camberlin, P., Janicot, S., & Poccard, I. (2001). Seasonality and atmospheric dynamics of
the teleconnection between African rainfall and tropical sea-surface temperature:
Atlantic vs. ENSO. International Journal of Climatology, 21, 973–1005.
Caminade, C., & Terray, L. (2010). Twentieth century Sahel rainfall variability as
simulated by the ARPEGE AGCM, and future changes. Climate Dynamics, 35, 75–94.
Deser, C., & Blackmon, M. L. (1993). Surface climate variations over the North Atlantic
Ocean during winter: 1900–1989. American Meteorological Society, 1743–1753.
Eklundh, L., & Olsson, L. (2003). Vegetation index trends for the African Sahel 1982–
1999. Geophysical Research Letters, 30.
Fensholt, R., & Rasmussen, K. (2011). Analysis of trends in the Sahelian ‘rain-use
efficiency’ using GIMMS NDVI, RFE and GPCP rainfall data. Remote Sensing of
Environment, 115, 438–451.
Folland, C. K., Palmer, T. N., & Parker, D. E. (1986). Sahel rainfall and worldwide sea
temperatures, 1901–85. Nature, 320, 602–607.
Giannini, A., Saravanan, R., & Chang, P. (2003). Oceanic forcing of Sahel rainfall on
interannual to interdecadal time scales. Science, 302, 1027–1030.
Hagos, S. M., & Cook, K. H. (2008). Ocean warming and late-twientieth-century Sahel
drought and recovery. Journal of Climate, 21, 3797–3814.
Heinemann, B.Open University Course Team. (1989). Ocean circulation. Oxford, U.K.
Pergamon Press.
Herman, A., Kumar, V. B., Arkin, P. A., & Kousky, J. V. (1997). Objectively determined 10-
day African rainfall estimates created for famine early warning systems.
International Journal of Remote Sensing, 18, 2147–2159.
Herrmann, S. M., Anyamba, A., & Tucker, C. J. (2005). Recent trends in vegetation
dynamics in the African Sahel and their relationship to climate. Global Environ-
mental Change-Human and Policy Dimensions, 15, 394–404.
Heumann, B. W., Seaquist, J. W., Eklundh, L., & Jonsson, P. (2007). AVHRR derived
phenological change in the Sahel and Soudan, Africa, 1982–2005. Remote Sensing of
Environment, 108, 385–392.
Hickler, T., Eklundh, L., Seaquist, J. W., Smith, B., Ardo, J., Olsson, L., Sykes, M. T., &
Sjostrom, M. (2005). Precipitation controls Sahel greening trend. Geophysical
Research Letters, 32.
Holben, B. N. (1986). Characteristics of maximum-value composite images from
temporal AVHRR data. International Journal of Remote Sensing, 7, 1417–1434.
Huber, S., Fensholt, R., & Rasmussen, K. (2011). Water availability as the driver of
vegetation dynamics in the African Sahel from 1982–2007. Global and Planetary
Change, 76, 186–195.
James, M. E., & Kalluri, S. N. V. (1994). The Pathfinder AVHRR land data set — An
improved coarse resolution data set for terrestrial monitoring. International Journal
of Remote Sensing, 15, 3347–3363.
Janicot, S., Harzallah, A., Fontaine, B., & Moron, V. (1998). West African monsoon
dynamics and eastern equatorial Atlantic and Pacific SST anomalies (1970–88).
Journal of Climate, 11, 1874–1882.
Jarlan, L., Tourre, Y. M., Mougin, E., Philippon, N., & Mazzega, P. (2005). Dominant
patterns of AVHRR NDVI interannual variability over the Sahel and linkages with
key climate signals (1982–2003). Geophysical Research Letters, 32.
Jury, M. R., & Mpeta, E. J. (2009). African climate variability in the satellite era.
Theoretical and Applied Climatology, 98, 279–291.
Lamb, P. J. (1978). Large-scale tropical Atlantic surface circulation patterns associated
with sub-Saharan weather anomalies. Tellus, 30, 240–251.
Le Houérou, H. N. (1989). The grazing land ecosystems of the African Sahel. Berlin; New
York: Springer-Verlag.
Lebel, T., & Ali, A. (2009). Recent trends in the Central and Western Sahel rainfall regime
(1990–2007). Journal of Hydrology, 375, 52–64.
Los, S. O. (1998). Estimation of the ratio of sensor degradation between NOAA AVHRR
channels 1 and 2 from monthly NDVI composites. IEEE Transactions on Geoscience
and Remote Sensing, 36, 206–213.
Lu, J. (2009). The dynamics of the Indian Ocean sea surface temperature forcing of Sahel
drought. Climate Dynamics, 33, 445–460.
Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M., & Francis, R. C. (1997). A Pacific
interdecadal climate oscillation with impacts on salmon production. Bulletin of the
American Meteorological Society, 78, 1069–1079.
Mohino, E., Janicot, S., & Bader, J. (2010). Sahel rainfall and decadal to multi-decadal sea
surface temperature variability. Climate Dynamics, 1–22.
Moron, V. (1994). Guinean and Sahelian rainfall anomaly indexes at annual and
monthly scales (1933–1990). International Journal of Climatology, 14, 325–341.
Nicholson, S. E., Davenport, M. L., & Malo, A. R. (1990). A comparison of the vegetation
response to rainfall in the Sahel and East-Africa, using Normalized Difference
Vegetation Index from NOAA AVHRR. Climatic Change, 17, 209–241.
Nicholson, S. E., Some, B., & Kone, B. (2000). An analysis of recent rainfall conditions in
West Africa, including the rainy seasons of the 1997 El Nino and the 1998 La Nina
years. Journal of Climate, 13, 2628–2640.
Oba, G., Post, E., & Stenseth, N. C. (2001). Sub-saharan desertification and productivity
are linked to hemispheric climate variability. Global Change Biology, 7, 241–246.
Palmer, T. N. (1986). Influence of the Atlantic, Pacific and Indian Oceans on Sahel
rainfall. Nature, 322, 251–253.
Philippon, N., Jarlan, L., Martiny, N., Camberlin, P., & Mougin, E. (2007). Characterization
of the interannual and intraseasonal variability of West African vegetation between
1982 and 2002 by means of NOAA AVHRR NDVI data. Journal of Climate, 20,
1202–1218.
Pinzon, J. E., Brown, M. E., & Tucker, C. J. (2005). Hilbert–Huang transform and its
applications. In N. E. Huang (Ed.), Interdisciplinary mathematical sciences, v. 5, (pp.
xii, 311 p.). Singapore; Hackensack, NJ; London: World Scientific.
Plisnier, P. D., Serneels, S., & Lambin, E. F. (2000). Impact of ENSO on East African
ecosystems: a multivariate analysis based on climate and remote sensing data.
Global Ecology and Biogeography, 9, 481–497.
Prince, S. D. (1991). A model of regional primary production for use with coarse
resolution satellite data. International Journal of Remote Sensing, 12, 1313–1330.
Prince, S. D. (1991). Satellite remote-sensing of primary production — Comparison of
results for Sahelian grasslands 1981–1988. International Journal of Remote Sensing,
12, 1301–1311.
Prince, S. D., De Colstoun, E. B., & Kravitz, L. L. (1998). Evidence from rain-use
efficiencies does not indicate extensive Sahelian desertification. Global Change
Biology, 4, 359–374.
Propastin, P., Fotso, L., & Kappas, M. (2010). Assessment of vegetation vulnerability to
ENSO warm events over Africa. International Journal of Applied Earth Observation
and Geoinformation, 12S, 83–89.
Raicich, F., Pinardi, N., & Navarra, A. (2003). Teleconnections between Indian monsoon
and Sahel rainfall and the Mediterranean. International Journal of Climatology, 23,
173–186.
Rasmussen, M. S. (1992). Assessment of millet yields and production in northern
Burkina Faso using integrated NDVI from the AVHRR. International Journal of
Remote Sensing, 13, 3431–3442.
Reynolds, R. W., Rayner, N. A., Smith, T. M., Stokes, D. C., & Wang, W. Q. (2002). An
improved in situ and satellite SST analysis for climate. Journal of Climate, 15,
1609–1625.
Rowell, D. P. (2003). The impact of Mediterranean SSTs on the Sahelian rainfall season.
Journal of Climate, 16, 849–862.
Saji, N. H., Goswami, B. N., Vinayachandran, P. N., & Yamagata, T. (1999). A dipole mode
in the tropical Indian Ocean. Nature, 401, 360–363.
Seaquist, J. W., Hickler, T., Eklundh, L., Ardo, J., & Heumann, B. (2008). Disentangling the
effects of climate and people on Sahel vegetation dynamics. Biogeosciences
Discussions, 5, 3045–3067.
Shanahan, T. M., Overpeck, J. T., Anchukaitis, K. J., Beck, J. W., Cole, J. E., Dettman, D. L.,
Peck, J. A., Scholz, C. A., & King, J. W. (2009). Atlantic forcing of persistent drought in
West Africa. Science, 324, 377–380.
Tucker, C. J., Pinzon, J. E., Brown, M. E., Slayback, D. A., Pak, E. W., Mahoney, R., Vermote,
E. F., & El Saleous, N. (2005). An extended AVHRR 8-km NDVI dataset compatible
with MODIS and SPOT vegetation NDVI data. International Journal of Remote
Sensing, 26, 4485–4498.
Tucker, C. J., Vanpraet, C. L., Sharman, M. J., & Vanittersum, G. (1985). Satellite remote-
sensing of total herbaceous biomass production in the Senegalese Sahel — 1980–
1984. Remote Sensing of Environment, 17, 233–249.
3284 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
UCAR (2011). http://www2.ucar.edu/news/backgrounders/arctic-oscillation-pineapple-
express-weather-maker-glossary (accessed Jan 2011)
Vermote, E., & Kaufman, Y. J. (1995). Absolute calibration of AVHRR visible and near-
infrared channels using ocean and cloud views. International Journal of Remote
Sensing, 16, 2317–2340.
Wang, G. L. (2003). Reassessing the impact of North Atlantic Oscillation on the sub-
Saharan vegetation productivity. Global Change Biology, 9, 493–499.
Ward,M. N. (1998). Diagnosis and short-lead time prediction ofsummerrainfall intropical
North Africa at interannual and multidecadal timescales. Journal of Climate, 11, 3167.
Williams, C. A., & Hanan, N. P. (2011). ENSO and IOD teleconnections for African
ecosystems: evidence of destructive interference between climate oscillations.
Biogeosciences, 8, 27–40.
Wolter, K., & Timlin, M. S. (1998). Measuring the strength of ENSO events: How does
1997/98 rank? Weather, 53, 315–324.
3285S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285

More Related Content

What's hot

Ecological and socio economic vulnerability to Climate change
Ecological and socio economic vulnerability to Climate changeEcological and socio economic vulnerability to Climate change
Ecological and socio economic vulnerability to Climate changemehebubsahana
 
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_California
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_CaliforniaCollins_et_al_-_Monitoring_shallow_landslides_in_SFB_California
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_CaliforniaMichael Walker Whitman
 
Planning For Climate Change In The Technical Analysis 6 9 09
Planning For Climate Change In The Technical Analysis 6 9 09Planning For Climate Change In The Technical Analysis 6 9 09
Planning For Climate Change In The Technical Analysis 6 9 09Michael DePue
 
Regional synthesis - Central America/Caribbean
Regional synthesis - Central America/CaribbeanRegional synthesis - Central America/Caribbean
Regional synthesis - Central America/Caribbeanipcc-media
 
Regional Climate Information: Small Islands - Tropical cyclones
Regional Climate Information: Small Islands - Tropical cyclones Regional Climate Information: Small Islands - Tropical cyclones
Regional Climate Information: Small Islands - Tropical cyclones ipcc-media
 
UndergradResearch
UndergradResearchUndergradResearch
UndergradResearchLaura Rook
 
Regional Climate Information: Small Islands - Pacific
Regional Climate Information: Small Islands - Pacific Regional Climate Information: Small Islands - Pacific
Regional Climate Information: Small Islands - Pacific ipcc-media
 
Tobias_Climate_Analysis_Final_Present
Tobias_Climate_Analysis_Final_PresentTobias_Climate_Analysis_Final_Present
Tobias_Climate_Analysis_Final_PresentJoshua Tobias
 
Modification and Climate Change Analysis of surrounding Environment using Rem...
Modification and Climate Change Analysis of surrounding Environment using Rem...Modification and Climate Change Analysis of surrounding Environment using Rem...
Modification and Climate Change Analysis of surrounding Environment using Rem...iosrjce
 
Global warming &climate changes
Global warming &climate changesGlobal warming &climate changes
Global warming &climate changesDr. sreeremya S
 
PringleStretch&Bardossy_NHESS2014
PringleStretch&Bardossy_NHESS2014PringleStretch&Bardossy_NHESS2014
PringleStretch&Bardossy_NHESS2014Justin Pringle
 
Saltwater intrusion model pompano fl
Saltwater intrusion model pompano flSaltwater intrusion model pompano fl
Saltwater intrusion model pompano flDarrel Dunn
 
Summary for Policy Makers
Summary for Policy MakersSummary for Policy Makers
Summary for Policy Makerscenafrica
 
Climate Change: Addressing the Major Skeptic Arguments
Climate Change: Addressing the Major Skeptic ArgumentsClimate Change: Addressing the Major Skeptic Arguments
Climate Change: Addressing the Major Skeptic ArgumentsAlexander Ainslie
 
Mekonnen adnew
Mekonnen adnewMekonnen adnew
Mekonnen adnewClimDev15
 
Presentation by Dr Omar Munyaneza
Presentation by Dr Omar Munyaneza Presentation by Dr Omar Munyaneza
Presentation by Dr Omar Munyaneza Bosco Hitimana
 
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSION
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSIONWE1.L09 - AN OVERVIEW OF THE DESDYNI MISSION
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSIONgrssieee
 
Regional Climate Information: Small Islands - Regional sea level rise and oce...
Regional Climate Information: Small Islands - Regional sea level rise and oce...Regional Climate Information: Small Islands - Regional sea level rise and oce...
Regional Climate Information: Small Islands - Regional sea level rise and oce...ipcc-media
 

What's hot (20)

Ecological and socio economic vulnerability to Climate change
Ecological and socio economic vulnerability to Climate changeEcological and socio economic vulnerability to Climate change
Ecological and socio economic vulnerability to Climate change
 
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_California
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_CaliforniaCollins_et_al_-_Monitoring_shallow_landslides_in_SFB_California
Collins_et_al_-_Monitoring_shallow_landslides_in_SFB_California
 
Planning For Climate Change In The Technical Analysis 6 9 09
Planning For Climate Change In The Technical Analysis 6 9 09Planning For Climate Change In The Technical Analysis 6 9 09
Planning For Climate Change In The Technical Analysis 6 9 09
 
Regional synthesis - Central America/Caribbean
Regional synthesis - Central America/CaribbeanRegional synthesis - Central America/Caribbean
Regional synthesis - Central America/Caribbean
 
Regional Climate Information: Small Islands - Tropical cyclones
Regional Climate Information: Small Islands - Tropical cyclones Regional Climate Information: Small Islands - Tropical cyclones
Regional Climate Information: Small Islands - Tropical cyclones
 
UndergradResearch
UndergradResearchUndergradResearch
UndergradResearch
 
Regional Climate Information: Small Islands - Pacific
Regional Climate Information: Small Islands - Pacific Regional Climate Information: Small Islands - Pacific
Regional Climate Information: Small Islands - Pacific
 
Tobias_Climate_Analysis_Final_Present
Tobias_Climate_Analysis_Final_PresentTobias_Climate_Analysis_Final_Present
Tobias_Climate_Analysis_Final_Present
 
AAG Tampa 2014
AAG Tampa 2014AAG Tampa 2014
AAG Tampa 2014
 
Modification and Climate Change Analysis of surrounding Environment using Rem...
Modification and Climate Change Analysis of surrounding Environment using Rem...Modification and Climate Change Analysis of surrounding Environment using Rem...
Modification and Climate Change Analysis of surrounding Environment using Rem...
 
Global warming &climate changes
Global warming &climate changesGlobal warming &climate changes
Global warming &climate changes
 
PringleStretch&Bardossy_NHESS2014
PringleStretch&Bardossy_NHESS2014PringleStretch&Bardossy_NHESS2014
PringleStretch&Bardossy_NHESS2014
 
Saltwater intrusion model pompano fl
Saltwater intrusion model pompano flSaltwater intrusion model pompano fl
Saltwater intrusion model pompano fl
 
Summary for Policy Makers
Summary for Policy MakersSummary for Policy Makers
Summary for Policy Makers
 
Climate Change: Addressing the Major Skeptic Arguments
Climate Change: Addressing the Major Skeptic ArgumentsClimate Change: Addressing the Major Skeptic Arguments
Climate Change: Addressing the Major Skeptic Arguments
 
ICES-ASC2010-v9
ICES-ASC2010-v9ICES-ASC2010-v9
ICES-ASC2010-v9
 
Mekonnen adnew
Mekonnen adnewMekonnen adnew
Mekonnen adnew
 
Presentation by Dr Omar Munyaneza
Presentation by Dr Omar Munyaneza Presentation by Dr Omar Munyaneza
Presentation by Dr Omar Munyaneza
 
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSION
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSIONWE1.L09 - AN OVERVIEW OF THE DESDYNI MISSION
WE1.L09 - AN OVERVIEW OF THE DESDYNI MISSION
 
Regional Climate Information: Small Islands - Regional sea level rise and oce...
Regional Climate Information: Small Islands - Regional sea level rise and oce...Regional Climate Information: Small Islands - Regional sea level rise and oce...
Regional Climate Information: Small Islands - Regional sea level rise and oce...
 

Similar to huber_2011_RSE

Forecasting monthly water resources conditions by using different indices
Forecasting monthly water resources conditions by using different indicesForecasting monthly water resources conditions by using different indices
Forecasting monthly water resources conditions by using different indicesAI Publications
 
Identification of three_dominant_rainfall_regions_
Identification of three_dominant_rainfall_regions_Identification of three_dominant_rainfall_regions_
Identification of three_dominant_rainfall_regions_Lasriama Siahaan
 
Tesfaye Samuel Presentation- MERGED.pptx
Tesfaye Samuel Presentation- MERGED.pptxTesfaye Samuel Presentation- MERGED.pptx
Tesfaye Samuel Presentation- MERGED.pptxTesfaye Samuel
 
Lesson6 greeningarctic20110315(2)
Lesson6 greeningarctic20110315(2)Lesson6 greeningarctic20110315(2)
Lesson6 greeningarctic20110315(2)Edie Barbour
 
Climate Change and Vegetation
Climate Change and VegetationClimate Change and Vegetation
Climate Change and VegetationGDCKUL
 
diurnal temperature range trend over North Carolina and the associated mechan...
diurnal temperature range trend over North Carolina and the associated mechan...diurnal temperature range trend over North Carolina and the associated mechan...
diurnal temperature range trend over North Carolina and the associated mechan...Sayem Zaman, Ph.D, PE.
 
Islam Intel poster FINAL
Islam Intel poster FINALIslam Intel poster FINAL
Islam Intel poster FINALTahsina Islam
 
turner_capstone_paper_042815 (1)
turner_capstone_paper_042815 (1)turner_capstone_paper_042815 (1)
turner_capstone_paper_042815 (1)Andre Turner
 
A high resolution-history_of_the_south_a
A high resolution-history_of_the_south_aA high resolution-history_of_the_south_a
A high resolution-history_of_the_south_aGeorgeaMelo1
 
ChapterClimate Change 2014Synthesis Report Summary.docx
ChapterClimate Change 2014Synthesis Report Summary.docxChapterClimate Change 2014Synthesis Report Summary.docx
ChapterClimate Change 2014Synthesis Report Summary.docxtiffanyd4
 
Hidrodinamica Cardoso&motta marques 2009aeco
Hidrodinamica Cardoso&motta marques 2009aecoHidrodinamica Cardoso&motta marques 2009aeco
Hidrodinamica Cardoso&motta marques 2009aecoCaline Gally
 
17 . 2 Climate Change In The Aral Sea Basin A Multi-Scale And Multi-Dimensi...
17 . 2 Climate Change In The Aral Sea Basin   A Multi-Scale And Multi-Dimensi...17 . 2 Climate Change In The Aral Sea Basin   A Multi-Scale And Multi-Dimensi...
17 . 2 Climate Change In The Aral Sea Basin A Multi-Scale And Multi-Dimensi...Jackie Gold
 
Historical rainfallvariabilitypaper
Historical rainfallvariabilitypaperHistorical rainfallvariabilitypaper
Historical rainfallvariabilitypaperPieterSteenkamp10
 
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Meyre Da Silva
 
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Da Silva Meyre
 
2011.12.16.UofMiami_Shukla.ppt
2011.12.16.UofMiami_Shukla.ppt2011.12.16.UofMiami_Shukla.ppt
2011.12.16.UofMiami_Shukla.pptStephenMcIntyre17
 

Similar to huber_2011_RSE (20)

Forecasting monthly water resources conditions by using different indices
Forecasting monthly water resources conditions by using different indicesForecasting monthly water resources conditions by using different indices
Forecasting monthly water resources conditions by using different indices
 
Identification of three_dominant_rainfall_regions_
Identification of three_dominant_rainfall_regions_Identification of three_dominant_rainfall_regions_
Identification of three_dominant_rainfall_regions_
 
#1
#1#1
#1
 
Tesfaye Samuel Presentation- MERGED.pptx
Tesfaye Samuel Presentation- MERGED.pptxTesfaye Samuel Presentation- MERGED.pptx
Tesfaye Samuel Presentation- MERGED.pptx
 
Lesson6 greeningarctic20110315(2)
Lesson6 greeningarctic20110315(2)Lesson6 greeningarctic20110315(2)
Lesson6 greeningarctic20110315(2)
 
Climate Change and Vegetation
Climate Change and VegetationClimate Change and Vegetation
Climate Change and Vegetation
 
diurnal temperature range trend over North Carolina and the associated mechan...
diurnal temperature range trend over North Carolina and the associated mechan...diurnal temperature range trend over North Carolina and the associated mechan...
diurnal temperature range trend over North Carolina and the associated mechan...
 
RGV Study Report
RGV Study ReportRGV Study Report
RGV Study Report
 
Islam Intel poster FINAL
Islam Intel poster FINALIslam Intel poster FINAL
Islam Intel poster FINAL
 
SterliniEtAl2015
SterliniEtAl2015SterliniEtAl2015
SterliniEtAl2015
 
5_Jerker et al_HESS_2012
5_Jerker et al_HESS_20125_Jerker et al_HESS_2012
5_Jerker et al_HESS_2012
 
turner_capstone_paper_042815 (1)
turner_capstone_paper_042815 (1)turner_capstone_paper_042815 (1)
turner_capstone_paper_042815 (1)
 
A high resolution-history_of_the_south_a
A high resolution-history_of_the_south_aA high resolution-history_of_the_south_a
A high resolution-history_of_the_south_a
 
ChapterClimate Change 2014Synthesis Report Summary.docx
ChapterClimate Change 2014Synthesis Report Summary.docxChapterClimate Change 2014Synthesis Report Summary.docx
ChapterClimate Change 2014Synthesis Report Summary.docx
 
Hidrodinamica Cardoso&motta marques 2009aeco
Hidrodinamica Cardoso&motta marques 2009aecoHidrodinamica Cardoso&motta marques 2009aeco
Hidrodinamica Cardoso&motta marques 2009aeco
 
17 . 2 Climate Change In The Aral Sea Basin A Multi-Scale And Multi-Dimensi...
17 . 2 Climate Change In The Aral Sea Basin   A Multi-Scale And Multi-Dimensi...17 . 2 Climate Change In The Aral Sea Basin   A Multi-Scale And Multi-Dimensi...
17 . 2 Climate Change In The Aral Sea Basin A Multi-Scale And Multi-Dimensi...
 
Historical rainfallvariabilitypaper
Historical rainfallvariabilitypaperHistorical rainfallvariabilitypaper
Historical rainfallvariabilitypaper
 
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
 
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
Physical Forcing Mechanisms Controlling the Variability of Chlorophyll-a over...
 
2011.12.16.UofMiami_Shukla.ppt
2011.12.16.UofMiami_Shukla.ppt2011.12.16.UofMiami_Shukla.ppt
2011.12.16.UofMiami_Shukla.ppt
 

huber_2011_RSE

  • 1. Analysis of teleconnections between AVHRR-based sea surface temperature and vegetation productivity in the semi-arid Sahel Silvia Huber ⁎, Rasmus Fensholt Dept. of Geography and Geology, University of Copenhagen, Øster Voldgade 10, DK-1350 Copenhagen, Denmark a b s t r a c ta r t i c l e i n f o Article history: Received 15 February 2011 Received in revised form 11 July 2011 Accepted 13 July 2011 Available online 13 September 2011 Keywords: Climate indices NDVI Oceanic forcing Remote sensing Sahel SST Vegetation dynamics Vegetation productivity across the Sahel is known to be affected by a variety of global sea surface temperature (SST) patterns. Often climate indices are used to relate Sahelian vegetation variability to large-scale ocean– atmosphere phenomena. However, previous research findings reporting on the Sahelian vegetation response to climate indices have been inconsistent and contradictory, which could partly be caused by the variations in spatial extent/definitions of climate indices and size of the region studied. The aim of this study was to analyze the linkage between climate indices, pixel-wise spatio-temporal patterns of global sea surface temperature and the Sahelian vegetation dynamics for 1982–2007. We stratified the Sahel into five subregions to account for the longitudinal variability in rainfall. We found significant correlations between climate indices and the Normalized Difference Vegetation Index (NDVI) in the Sahel, however with different magnitudes in terms of strength for the western, central and eastern Sahel. Also the correlations based on NDVI and global SST anomalies revealed the same East–West gradient, with a stronger association for the western than the eastern Sahel. Warmer than average SSTs throughout the Mediterranean basin seem to be associated with enhanced greenness over the central Sahel whereas colder than average SSTs in the Pacific and warmer than average SSTs in the eastern Atlantic were related to increased greenness in the most western Sahel. Accordingly, we achieved high correlations for SSTs of oceanic basins which are geographically associated to the climate indices yet by far not always these patterns were coherent. The detected SST–NDVI patterns could provide the basis to develop new means for improved forecasts in particular of the western Sahelian vegetation productivity. © 2011 Elsevier Inc. All rights reserved. 1. Introduction The development and prosperity of the African Sahel, a semi-arid transition zone between the Sahara desert to the north and the humid tropical savanna to the south, largely depend on the rainfall regime. Vegetation growth is limited to the short rainy season with July to September being the months of highest rainfall (Nicholson et al., 2000). The vegetation phenology thereby closely responds to the seasonal cycle of rain with vegetation productivity taking place mainly during the months with rainfall (Herrmann et al., 2005; Hickler et al., 2005; Huber et al., 2011; Prince et al., 1998). However, precipitation in the Sahel shows typically large interannual and decadal variability. This variability was impressively manifested starting in the mid 1960s, when rainfall amounts continuously declined across the Sahel culminating in a widespread drought in 1982–1985. But since then, rainfall has steadily increased together with the vegetation resource base. This greening up has been observed from satellite based Normalized Difference Vegetation Index (NDVI) time series (Anyamba and Tucker, 2005; Eklundh and Olsson, 2003; Herrmann et al., 2005; Heumann et al., 2007; Seaquist et al., 2008). Due to the close coupling between rainfall and vegetation growth in the Sahel, rainfall variability explains a significant part of this increase in NDVI (Fensholt and Rasmussen, 2011; Herrmann et al., 2005; Hickler et al., 2005;Huber et al., 2011;Nicholsonetal., 1990).NDVI has,despite of the index simplicity, shown to be a reliable proxy for vegetation vigor and productivity in the Sahel (Prince, 1991a,b; Rasmussen, 1992; Tuckeret al., 1985). The reasons for the large interannual and decadal fluctuations in rainfall are still not entirely clear but the early works of Folland et al. (1986), Lamb (1978) and Palmer (1986) highlighted that rainfall variability across the Sahel is related (teleconnected) with regional and global sea surface temperature (SST) conditions. These include interannual and decadal temperature anomalies in the Atlantic (e.g., Biasutti et al., 2008; Palmer, 1986; Shanahan et al., 2009; Ward, 1998), the Pacific (e.g., Caminade and Terray,2010; Janicot et al., 1998; Mohino et al., 2010), the Indian Ocean (e.g., Bader and Latif, 2003; Giannini et al., 2003; Lu, 2009) and the Mediterranean (e.g., Philippon et al., 2007; Raicich et al., 2003; Rowell, 2003). Often climate indices are used to relate Sahelian rainfall or vegetation variability to large-scale ocean– atmosphere phenomena. Climate indices reflect the essential elements Remote Sensing of Environment 115 (2011) 3276–3285 ⁎ Corresponding author. Tel.: +45 35322 499. E-mail addresses: huber.silvia@gmail.com (S. Huber), rf@geo.ku.dk (R. Fensholt). 0034-4257/$ – see front matter © 2011 Elsevier Inc. All rights reserved. doi:10.1016/j.rse.2011.07.011 Contents lists available at SciVerse ScienceDirect Remote Sensing of Environment journal homepage: www.elsevier.com/locate/rse
  • 2. of climate and its fluctuations through time. They can either be very simple, for example only calculated from SSTs, or integrate a larger number of climate variables, such as temperature, pressure or surface winds. Some relate to interannual climate patterns, while others reflect decadal to inter-decadal phenomena. The El Niño-Southern Oscillation (ENSO) is one of the best known indices of interannual climate variability. However, its impact on the Sahelian vegetation is still debated. While several studies have reported no or only a minor influence (Anyamba et al., 2001; Anyamba and Eastman, 1996; Anyamba and Tucker, 2005; Philippon et al., 2007; Propastin et al., 2010), others have shown that the Sahelian climate is related to ENSO events (Camberlin et al., 2001; Oba et al., 2001; Ward, 1998). Other climate indices that were associated to the Sahelian vegetation dynamics are the North Atlantic Oscillation (NAO), the Pacific Decadal Oscillation (PDO) and the Indian Ocean Dipole (IOD). Oba et al. (2001) attributed in their study large parts of the interannual variation of vegetation productivity during the 1980s to the NAO while Wang (2003) could not find a consistent relationship. The influence of the PDO and the IOD on the Sahelian climate is less studied. Brown et al. (2010) found significant relationships between the PDO and the growing season in West Africa while the IOD was found to have virtually no influence in the Sahel. Also Williams and Hanan (2011) found only weak responses across the Sahelian zone in their study when looking at the IOD and the Multivariate ENSO Index (MEI) individually but not when investigating interacting effects of the two indices. All these findings suggest some contradiction in the Sahelian response to changes as characterized by climate indices, which could partly be caused by the variations in spatial extent respectively definitions of climate indices and size of the region studied. The goal of this paper is to analyze the linkage between these climate indices, pixel-wise spatio-temporal patterns of global sea surface temperature and the Sahelian vegetation dynamics. In this analysis the Sahel, an ecoregion covering an area of more than 3 million km2 , was divided into five subregions because it has been shown that the Atlantic section of the Sahel is only weakly related with the rest of the zone (Moron, 1994). Moreover the definition of the “Sahel window” varies considerably among different studies, in particular in the longitudinal extent. For example Giannini et al. (2003) referred to the 20°W–35°E/10–20°N extent, Mohino et al. (2010) used the area 15°W–15°E/10–17°N, Jarlan et al. (2005) studied the window 18°W–18°E/12–18°N and Rowell (2003) defined the area 16.875°W–35.625°E/11.25–18.75°N as the Sahel. Here we used the longitudinal extent defined in the study of Giannini et al. (2003) (20°W–35°E) but stratified the Sahel into five subregions as displayed in Fig. 1 to account for the longitudinal variability in rainfall. Using Pearson's correlation analysis between climate indices, remotely sensed time series of SST and NDVI corrected for seasonality and trend, the following research questions were investigated: a) is there a statistically significant relationship between global SST anomalies/climate indices and NDVI in the Sahel and are the spatial patterns consistent; b) which ocean regions explain most of the NDVI dynamics across the subdivided Sahel; c) which time lags are involved in the interrelations between SST/climate indices and NDVI and d) do certain areas in the Sahel exhibit stronger teleconnections to SST anomalies than others. The spatio-temporal correlation analysis also including sea surface temperature may help to unveil new options to develop tools for forecasting the Sahelian resource base, which is largely dependent on rainfall. 2. Data Four climate indices and two different satellite-based gridded monthly time series (NDVI and SST) were used in this study. High- temporal continuous EO-based NDVI and SST is currently available from 1982 to 2007 (full year coverage=312 months). This time period thus determines the temporal extent of the analyses performed. 2.1. Climate indices We used four climate indices based on global climate variations from various oceanic regions (Fig. 2). 2.1.1. Indian Ocean Dipole (IOD) The Indian Ocean Dipole (IOD) is an interannual climate phenomenon in the tropical Indian Ocean, first defined by Saji et al. (1999). In the positive phase of the IOD, trade winds are stronger than usual and cooler-than-average sea-surface temperatures are preva- lent across the eastern tropical Indian Ocean, near Indonesia and Australia. To the west, near Madagascar, waters are warmer than average and convection is intensified. These patterns are reversed during the IOD's negative phase (UCAR, 2011). Intensity of the IOD is represented by anomalous SST gradients between the western equatorial Indian Ocean (50E–70E and 10S–10N) and the south eastern equatorial Indian Ocean (90E–110E and 10S–0N). We used the IOD based on HadlSST SST (monthly from 1982 to 2007) downloaded from: http://www.jamstec.go.jp/frsgc/research/d1/iod/. 2.1.2. Multivariate ENSO Index (MEI) The Multivariate El Niño-Southern Oscillation (ENSO) Index (MEI) is a measure to monitor the strength of ENSO conditions. ENSO is the most important coupled ocean–atmosphere phenomenon to cause global climate variability on interannual time scales. MEI is derived by combining the six main observed variables in the tropical Pacific: SST, sea-level pressure, surface winds, surface air temperature, cloudiness and precipitation (Wolter and Timlin, 1998). The MEI key regions for the variable measurements can be seen in Wolter and Timlin (1998, Fig. 1). MEI is favored over conventional indices, since it combines the significant features of all observed surface parameters in the tropical Pacific (Wolter and Timlin, 1998). MEI was obtained from: http:// www.esrl.noaa.gov/psd/data/climateindices/. 2.1.3. North Atlantic Oscillation Index (NAO) The North Atlantic Oscillation (NAO) is typically measured through variations in the normal pattern of lower atmospheric Fig. 1. Map of the average annual rainfall in the Sahel (1996–2007) and the areas of study: two regions have been selected in the western (W1, W2) and central Sahel (C1, C2) and one region in the eastern Sahel (E1). 3277S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 3. pressure over Iceland and higher pressure near the Azores and Iberian Peninsula (UCAR, 2011). A positive NAO index refers to stronger than usual subtropical high pressure center around the Azores and a deeper than normal Icelandic low, with increased pressure generating more and stronger winter storms crossing the Atlantic Ocean on a more northerly track. Europe tends towards warm and wet winters while northern Canada and Greenland will usually have cold and dry winters, with the eastern United States generally experiencing mild, wet winter conditions. A negative NAO index (when there is less difference than usual in pressure across the two regions) features a weakened Atlantic storm track, a greater risk for Arctic outbreaks of cold air across the northeastern United States and northern Europe, and moist air to the Mediterranean (UCAR, 2011). NAO varies from year to year but has a roughly decadal pattern with a dominant period of 12 years (Deser and Blackmon, 1993). The index can be obtained from: http://www. esrl.noaa.gov/psd/data/climateindices/. 2.1.4. Pacific Decadal Oscillation (PDO) The Pacific Decadal Oscillation (PDO) is a multi-decadal pattern of climate variability across the North Pacific Ocean (Mantua et al., 1997). The positive (warm) mode of PDO features colder than average SSTs in the central North Pacific along a narrow band of warmer SSTs along the west coast of North America and in the eastern tropical Pacific. During the negative (cool) phase of PDO, the opposite has been observed: a warm pool of sea surface waters in the central north Pacific and cold SSTs along the west coast (Mantua et al., 1997). Each phase typically persists for 20 to 30 years, with a warm phase predominating since the late 1970s. The PDO may be related to ENSO, but differs mainly because the timescale for the PDO is much longer (several decades) and because the PDO more clearly involves the extratropical Pacific and the Aleutian Low pressure system (UCAR, 2011). Even though PDO is mirroring decadal patterns, it involves sufficient interannual variability to relate to productivity in Africa (Brown et al., 2010). The index is available at http://jisao.washington. edu/pdo/PDO.latest. 2.2. Normalized Difference Vegetation Index (NDVI) We used AVHRR (Advanced Very High Resolution Radiometer) GIMMS (Global Inventory Modeling and Mapping Studies) NDVI data as a proxy for vegetation productivity. Monthly maximum NDVI composites with an 8 km spatial resolution used in this study were processed by the GIMMS Group at NASA's Goddard Space Flight Center, as described by Tucker et al. (2005) and Pinzon et al. (2005). More details about the binning and band calibration can be found in James and Kalluri (1994), Vermote and Kaufman (1995) and Los (1998). No atmospheric correction is applied to the GIMMS data except for volcanic stratospheric aerosol periods (1982–1984 and 1991–1994) (Tucker et al., 2005). A satellite orbital drift correction is performed using an empirical mode decomposition (EMD) transfor- mation method of Pinzon et al. (2005) removing common trends between time series of solar zenith angle (SZA) and NDVI. The original 16-bit GIMMS NDVI was converted into real NDVI values (range −1 to 1) for further analysis. 2.3. Sea surface temperature (SST) NOAA Optimum Interpolation (OI) SST v2 data, provided by the NOAA-CIRES Climate Diagnostics Center, Boulder, USA were used in this paper (http://www.esrl.noaa.gov/psd/data/gridded/data.noaa. oisst.v2.html). The NOAA SST OIv2 product integrates both in situ and satellite data from November 1981 to the present at 1.0° spatial resolution globally in degrees Celsius (Reynolds et al., 2002). The in situ SST data are determined from observations from ships and buoys (Reynolds et al., 2002). Satellite data is also obtained from the AVHRR instrument. The NOAA OI.v2 SST monthly fields are derived by a linear interpolation of the weekly optimum interpolation (OI) version 2 fields to daily fields and then averaging the daily values over a month. More details about the product can be found in Reynolds et al. (2002). 3. Methods The boundary of the Sahel used in this paper (indicated on the maps in Figs. 1, 4 and 5) is defined by Le Houérou (1989) using the IOD -3.5 -2.5 -1.5 -0.5 0.5 1.5 2.5 3.5 1982 1985 1988 1991 1994 1997 2000 2003 2006 Year 1982 1985 1988 1991 1994 1997 2000 2003 2006 Year 1982 1985 1988 1991 1994 1997 2000 2003 2006 Year 1982 1985 1988 1991 1994 1997 2000 2003 2006 Year NormalizedIndex -3.5 -2.5 -1.5 -0.5 0.5 1.5 2.5 3.5 NormalizedIndex -3.5 -2.5 -1.5 -0.5 0.5 1.5 2.5 3.5 NormalizedIndex -3.5 -2.5 -1.5 -0.5 0.5 1.5 2.5 3.5 NormalizedIndex MEI NAO PDO Fig. 2. Monthly climate indices Indian Ocean Dipole (IOD), Multivariate ENSO Index (MEI), North Atlantic Oscillation Index (NAO) and Pacific Decadal Oscillation (PDO) from 1982 to 2007. 3278 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 4. 150 (to the north) and 700 mm (to the south) annual rainfall isohyets as borders of the Sahelian area based on rainfall mean annual values of the Rainfall Estimate (RFE) blended gage-satellite rainfall product (NOAA Climate Prediction Center) for 1996–2007 (Herman et al., 1997). Within this boundary we defined five subregions (Fig. 1): two in the Western (W1, W2), two in the Central (C1, C2) and one in the Eastern Sahel (E1) to account for different rainfall regimes in the East– West orientation at the interannual scale. The longitudinal division was partly based on the analysis conducted by Moron (1994) and Lebel and Ali (2009). Three subregions (W1, W2 and C2) are 10° wide in longitude and C1 and E1 were chosen to be 15° wide in latitude. It was decided to define C1 with a larger width to keep Chad incl. the Marra mountains as well as the Sahelian part of Sudan as an entity. E1 was enlarged to 15° since rainfall anomalies of the most eastern Sahel (35–40°E) and the region between 25° and 35°E are highly correlated. 3.1. Preprocessing All analyses were performed on linearly detrended NDVI and SST time series. Further, for both time series standardized anomalies were computed to remove the seasonal component from the data. The original 15-day NDVI composite data were aggregated to months using a maximum value composite approach to further reduce the influence from clouds (Holben, 1986). Monthly anomalies for each 26-year data record (1982–2007=312 images) were obtained by computing the median value for each pixel for each month (=climatology value) which was then subtracted from each image. Median values rather than mean values were used because the time series of this study are shorter than the standard 30-year baseline period defined by the World Meteorological Organization (WMO) to calculate climatology values. The SST time series as well as the climate indices were additionally smoothed using a moving average filter over a 3-month period (Jan–March, Feb–April etc.) to reduce noise (Plisnier et al., 2000). From the NDVI time series mean values for the period July– September (termed JAS hereafter) were calculated for each of the five subregions in the Sahel. Finally, area-average mean JAS NDVI were extracted for all subregions, termed hereafter NDVI Anomaly Indices (NAI) (Fig. 3). 3.2. Teleconnections Pixel-wise Pearson correlation coefficients were calculated be- tween the four climate indices (IOD, MEI, NAO and PDO), global SST anomalies (3-month means) and NAI values for each of the five subregions from 1982 to 2007. Since changes in ocean-atmospheric patterns can affect the Sahelian rainfall regime and hence the vegetation development with a delay of several months, lagged correlation was used, with climate indices and SST observed at steps 0 to 9 months prior to the NDVI observations in the Sahel. Next, oceanic regions (all substantially larger than the Sahel) with highest positive or negative correlations with JAS NDVI were identified and three indices of SST anomalies extracted (Fig. 4). Two indices were extracted in the northern Atlantic Ocean. The first one for January– March mean SSTs as displayed in Fig. 6b: JFM) (N3.5 million km2 ) and the second one for June–August mean SSTs as illustrated in Fig. 6a: JJA) (N4.5 million km2 ). The third index covers an area of more than 5 million km2 in the equatorial Pacific and was extracted from March– May mean SSTs (Fig. 6a: MAM). These three SST indices (JFM Atlantic, JJA Atlantic and MAM pacific) were correlated with JAS NDVI across the Sahel and the results compared with the commonly used climate indices. To explore combined effects of the indices, partial correlation analysis was applied. 4. Results and discussion Correlations of JAS NDVI anomalies (NAI) between the different sub-regions of the Sahel confirm that the NDVI dynamics of the Atlantic section (W1) is to a certain degree decoupled from the rest of the region (Table 1) as also suggested by (Moron, 1994). The subregion W2 is closer related to the central Sahel (C1) than to W1. The highest correlation was found between W2 and C1 (r=0.78). 4.1. Correlations between climate indices and Sahelian NDVI dynamics The analysis revealed that the western and central Sahel is better correlated to predefined climate indices (IOD, MEI, NAO and PDO) than the eastern Sahel. Most significant correlations were achieved for MEI and PDO (Fig. 5a and b). In particular for Senegal, southern Mauritania and western Mali we found highly significant r-values for MEI with the highest number of significant pixels for the periods AMJ and MJJ. Brown et al. (2010) also reported significant correlations between MEI aggregated over June to August (JJA) and the start of the growing season (SOS) for the westernmost Sahel, however in their study they report positive correlations while we found values of the opposite sign. Other studies report only modest to small ENSO response of NDVI for the Sahel (Anyamba and Eastman, 1996; Philippon et al., 2007; Williams and Hanan, 2011). Also for PDO significant negative correlations were found throughout the western to central Sahel; in particular areas of Mali Fig. 3. Area-averaged JAS NDVI standardized anomalies (NAI) obtained from linearly detrended time-series from 1982 to 2007 for the five Sahelian subregions as outlined in Fig. 1. 3279S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 5. and Burkina Faso were related to the index (Fig. 5b). The highest anticorrelations for PDO and JAS NAI were found when averaging PDO over JFM and JAS. In contrast to the results obtained for MEI and PDO, we found some positive correlations when testing associations between JAS NAI and IOD and NAO indices, respectively, (Fig. 5c and d); yet a smaller number of pixels with significant r-values was found as compared to MEI and PDO which is in accordance with the findings of Brown et al. (2010). 4.2. Correlations between global SST and Sahelian NDVI dynamics The pixel-wise correlation analysis revealed several significant relationships between SST and vegetation dynamics in the Sahel (Fig. 6). In general larger oceanic areas with significant correlations were found for the western and central Sahel than for the eastern Sahel. This is in accordance with the findings presented in the previous section. Subsequently we report teleconnections between JAS NAI and SST anomalies for the five different subregions. Only large scale patterns of significant correlations that are persistent over at least three aggregated time periods will be commented on. 4.2.1. Sahel subregion W1 (20–10° W) The results obtained for W1 differ from the other four subregions, corroborating the fact that the Atlantic section of the Sahel is less well related to the rest of the ecoregion as described above. Compared to the other subregions, W1 in general shows stronger associations to oceanic temperature anomalies reflected by more pixels with significant r-values (Fig. 6a). In particular large areas of the Pacific Ocean emerge with significant correlations (p≤0.01), both positive and negative. Anticorrelations are very distinct in the tropical Pacific, with decreasing r-values for shorter lag-periods while the areas to the north and south of the tropical Pacific exhibit positive correlations. For the Philippine Sea also positive relationships were found between SST and JAS NAI anomalies for the periods JFM, FMA and MAM. The anticorrelations imply that lower than average SSTs in the tropical Pacific Ocean are statistically related to higher boreal summer precipitation in the West Sahel (W1). These results are consistent with other research. Ward (1998) found similar patterns for “no dipole years”, i.e., years with rainfall anomalies of the same sign in the Sahel and the Guinean Coast at high-frequency timescale, when correlating JAS SST and Sahelian rainfall. Higher photosynthetic activity over the Sahel was also associated to negative summer SST anomalies in the tropical eastern Pacific in the study of Philippon et al. (2007). However, in this study we could link winter/ spring SSTs to summer NDVI in the Sahel. This time lag could be important for predictive purposes. The strong pattern we obtained in the Pacific Ocean would suggest that there is also a link between PDO and JAS NAI in the Western Sahel, but this was not the case (Fig. 5b). For shorter lag-periods (starting with MJJ) high anticorrelations emerged in the Indian and Southern Ocean, with r-values between − 0.60 and −0.68. Jury and Mpeta (2009) showed an association between the westerly wind anomalies and sea-level pressure of the South Indian Ocean and Sahel rainfall with a 12-month lead time. The Indian Ocean in general plays an important role in forcing the Sahelian climate, in particular at decadal time scales, as reported in various studies (e.g., Bader and Latif, 2003; Giannini et al., 2003; Hagos and Cook, 2008). Philippon et al. (2007) however did not find any significant signal over the tropical Indian Ocean. They argue in their study that this is merely attributed to the time period that is restricted to the recent two decades, which does not capture longterm decadal variability. Even though the time scale covered in this study does not allow detecting decadal or multi-decadal patterns, patterns in the Indian Ocean showed to be associated with JAS NAI. Together with the occurrence of significant patterns in the Indian Ocean, positive correlations occur in the western part of the North Atlantic Ocean, from the MJJ period onwards (p≤0.01, with highest correlations for JAS (r=0.84)) characterizing a large region of pixels (horseshoe shaped) stretching from just off the coast of Senegal to the Atlantic ocean west of the European continent. This spatial extent corresponds well with the cold ocean current associated with one of the five major existing gyres (the North Atlantic gyre) (Heinemann and Open University Course Team, 1989). However, even though we found these positive correlations in the northern Atlantic and a weak correlation (r=0.35) between JAS NAO and the JJA Atlantic index (extracted in this study and described in Section 3.2), almost no significant relationships were found when correlating NAO with JAS NAI for the subregion W1. 4.2.2. Sahel subregion W2 (10–0° W) For JAS NAI of W2 in particular two oceanic areas emerged with highly significant negative correlations (p≤0.01): an area of the Atlantic Ocean for longer lag-periods (JFM, FMA and MAM) and an area of the Indian Ocean with a maximum r value of −0.8 for SST anomalies averaged for MJJ (Fig. 6b). The strong link between northern Atlantic SSTs and NDVI anomalies in the subregion W2 is in agreement with the findings of the correlation analysis between JAS NAI and NAO. Also for NAO we found significant (but weak) correlations with W2 NDVI for long lag-periods, in particular JFM, Fig. 4. Extracted SST anomaly indices for the Atlantic averaged over January–March (JFM) and June–August (JJA), respectively, and the equatorial Pacific for March–May (MAM). Table 1 Pearson's correlation coefficients between area-averaged JAS NDVI anomaly indices (NAI) for the five subregions in the Sahel. Subregion W1 W2 C1 C2 E1 W1 [20–10°W] 0.44 0.17 0.12 0.03 W2 [10–0°W] 0.44 0.78 0.43 0.26 C1 [0–15°E] 0.17 0.78 0.63 0.35 C2 [15–25°E] 0.12 0.43 0.63 0.63 E1 [25–40°E] 0.03 0.26 0.35 0.63 3280 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 6. even though NAO represents pressures which are not directly comparable to SST patterns. Compared to the results of W1 a general decrease in the geographical extent of significant pixels can be observed. In particular the association to the Pacific Ocean is much weaker. For example, W2 is the only subregion for which almost no significant relationships were detected in the Western Pacific. But still we identified significant negative correlations between JAS NAI of this region (and also C1 and C2, but not for W1) and the PDO. Interestingly some SST patterns in the North Pacific and the Aleutian Islands temporally coincide with the strongest NAI–PDO relationships. PDO has been shown to clearly involve the extratropical Pacific and the Aleutian low pressure system (UCAR, 2011). Finally, while we detected for W1 strong positive correlations from MJJ onwards in the northern Atlantic, for W2 the analysis revealed negative correlations for JFM until AMJ. The reason for this pattern needs to be further investigated. 4.2.3. Sahel subregion C1 (0–15° E) Similar to W2, also for the JAS NDVI dynamics of C1 anticorrelations were discovered in the Indian Basin with max. r-values of −0.63 for the period MAM (Fig. 6c). Starting with AMJ the Eastern Mediterranean Sea showed increasing associations with NAI. The findings that the Mediterranean plays an important role in the Sahelian climate are consistentwiththoseof other studies (Raicich et al., 2003; Rowell, 2003) and suggest that a warming of the Mediterranean is often associated with enhanced Sahelian rainfall and hence increased vegetation growth. The main reason for the regional teleconnection leading to additional moisture over the Sahel in warm Mediterranean years is increased evaporation over the sea surface leading to an enhanced moisture content of the air that is advected southwards across the eastern Sahara into the Sahel (Rowell, 2003). During winter, the whole Mediterranean is influenced by westerlies (Raicich et al., 2003). During summer the westerlies are weaker and a meridional regime develops, especially over the eastern Mediterranean basin (=Etesian wind regime) which connects the eastern Mediterranean area with the sub-Saharan ecoregion (Raicich et al., 2003). For the period JFM and FMA the western Pacific also partly contributes to JAS NDVI variability. 4.2.4. Sahel subregion C2 (15–25° E) For C2 positive relationships were found between JAS NAI and SST anomalies in the South West pacific for longer lag-periods (JFM, FMA, MAM) (Fig. 6d). Only in July–September (JAS) the SST anomalies measured in the SE Mediterranean and the Red Sea corresponded well with NAI (r=0.65). As for C1, the study also revealed significant positive correlations for the western Pacific for JFM, FMA and MAM. 4.2.5. Sahel subregion E1 (25–40° E) With the JS NAI of this region the SST anomalies of only a few oceanic areas were significantly correlated (Fig. 6e). Teleconnections were found almost exclusively in the Atlantic Ocean and Western Pacific for longer lag-periods. The last two periods analyzed (JJA and Fig. 5. Maps of significant r-values from correlation analyses between JAS NDVI anomalies and a) the Multivariate ENSO Index (MEI) averaged over May–July, b) the Pacific Decadal Oscillation (PDO) averaged over July–September, c) the North Atlantic Oscillation (NAO) averaged over January–March and d) the Indian Ocean Dipole (IOD) averaged over April– June from 1982 to 2007. 3281S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 7. Fig. 6. Maps of correlation coefficients (r) between the Sahel NDVI anomaly index (NAI) extracted from the five subregions: a) W1, b) W2, c) C1, d) C2 and e) E1) and mean SST anomalies from 1982 to 2007 for different time lags (e.g., JFM: January–March). Only correlations between the 1 and 5% confidence levels are shown. 3282 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 8. JAS) revealed positive associations between the SST of the Red Sea and the NDVI dynamics of the Eastern Sahel as it was the case for the neighboring subregion C2. These findings correspond with the climate indices analysis, which revealed hardly any significant correlations for the most eastern Sahel. 4.3. Joint correlations between climate/SST indices and Sahelian NDVI dynamics The results obtained from a partial correlation analysis are presented in Fig. 7. It can be seen that in certain hotspot regions across the Sahel up to 50% of the interannual variability in NDVI can be explained by the combined indices. With the combined JFM Atlantic and MAM Pacific indices hotspots of high r2 -values were mapped across the Sahel, except for C2 (Fig. 7b). Interestingly, highest correlations were achieved for C2 with MJJ MEI and JAS PDO (Fig. 7a). When the JJA Atlantic index was used in combination with the MAM Pacific index, in particular large parts of the NDVI variability of the western Sahel can be explained (Fig. 7c). In Senegal, Mali and Mauritania large areas are mapped with r2 -values between 0.4 and 0.5. These results illustrate that a larger amount of the JAS NDVI variability in the western Sahel can be explained when using extracted SST indices as compared to the well known MEI and PDO. Yet the latter were able to explain more of the JAS NDVI variability in subregion C2. 5. Conclusions Using EO-based time series of SST and NDVI covering 1982–2007 provides the possibility of performing analyses based on observations continuous in space and time but on the other hand the time record of the AVHRR sensors is not long enough to discover decadal patterns. We therefore focused on interannual patterns in this study. From this analysis we conclude that significant correlations exist between global SST anomalies and Sahelian NDVI, however with different characteristics for western, central and eastern Sahel. Whereas the vegetation productivity in the western Sahel could be associated with large oceanic areas of the Pacific, the Atlantic as well as the Indian Ocean, for the eastern Sahel only small areas in the Atlantic were found to be significantly related to dynamics in NDVI. The Eastern Mediterranean emerged only with significant r-values when related to NDVI in the central Sahel. Warmer than average SSTs throughout the Mediterranean basin seem to be associated with enhanced greenness over the central Sahel whereas colder than average SSTs in the Pacific and warmer than average SSTs in the eastern Atlantic were related to increased greenness in the most western Sahel. The correlations based on NAI and climate indices revealed the same East–West gradient, with stronger associations for the western than the eastern Sahel. Accordingly, we achieved high correlations for SSTs of oceanic basins which are associated to the indices (e.g., for W1: MEI and equatorial Pacific) yet by far not always. For instance relating IOD with Sahelian NDVI did not result in higher correlations even though the SSTs of the Indian Ocean played an important role for the NDVI dynamics in W1, W2 and C1. This result may be explained by the fact that IOD is defined as a gradient in the equatorial Indian Ocean. However, for W2 and C1 the most significant correlations were found in the southern Indian Ocean, away from the Equator and for W1, the correlations are significant over the Equator but show the same signed structure and no gradient. For NDVI of W1 the study showed a strong association to the SST anomalies of the Pacific Ocean but when relating PDO to JAS NAI this link could not be reproduced for W1. Overall, these large scale climate indices do have predictive power but they might be defined too broad thereby suppressing predictive capabilities of more localized areas like it seems to be the case for the Atlantic SST anomaly we found along the Senegal–Europe area for JAS. This is reflected in the findings of the correlation analysis based on combined indices. It illustrates that with extracted SST indices for Fig. 7. Maps of joint correlations (r2 ) from partial correlation analysis of JAS NDVI anomalies and a) the Multivariate ENSO Index (MEI) averaged over May–June and the Pacific Decadal Oscillation (PDO) averaged over June–August, b) the SST indices extracted from the Atlantic and Pacific for January–March and March–May, respectively, and c) the SST indices extracted from the Atlantic and Pacific for June–August and March–May, respectively. 3283S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 9. specific oceanic areas the percentage of explained NDVI variability can be increased and extended to larger areas as compared to traditional climate indices. However, in this paper we only investigated linear SST–NDVI relationships. Also non-linear features or interferences between climate patterns might play a role and this should be considered in future studies. The detected SST–NDVI patterns could provide the basis to develop new means for improved forecasts in particular of the western Sahelian vegetation resource base for pastoralism and agricultural production. Acknowledgements The authors thank the NASA Global Inventory Modeling and Mapping Studies (GIMMS) group for producing and sharing the AVHRR GIMMS NDVI data set as well as the NOAA-CIRES Climate Diagnostics Center, Boulder, USA, for providing the NOAA Optimum Interpolation (OI) SST v2 data. Thanks to two anonymous reviewers for their helpful comments. References Anyamba, A., & Eastman, J. R. (1996). Interannual variability of NDVI over Africa and its relation to El Nino Southern Oscillation. International Journal of Remote Sensing, 17, 2533–2548. Anyamba, A., & Tucker, C. J. (2005). Analysis of Sahelian vegetation dynamics using NOAA-AVHRR NDVI data from 1981–2003. Journal of Arid Environments, 63, 596–614. Anyamba, A., Tucker, C. J., & Eastman, J. R. (2001). NDVI anomaly patterns over Africa during the 1997/98 ENSO warm event. International Journal of Remote Sensing, 22, 1847–1859. Bader, J., & Latif, M. (2003). The impact of decadal-scale Indian Ocean sea surface temperature anomalies on Sahelian rainfall and the North Atlantic Oscillation. Geophysical Research Letters, 30. Biasutti, M., Held, I. M., Sobel, A. H., & Giannini, A. (2008). SST forcings and Sahel rainfall variability in simulations of the twentieth and twenty-first centuries. Journal of Climate, 21, 3471–3486. Brown, M. E., de Beurs, K., & Vrieling, A. (2010). The response of African land surface phenology to large scale climate oscillations. Remote Sensing of Environment, 114, 2286–2296. Camberlin, P., Janicot, S., & Poccard, I. (2001). Seasonality and atmospheric dynamics of the teleconnection between African rainfall and tropical sea-surface temperature: Atlantic vs. ENSO. International Journal of Climatology, 21, 973–1005. Caminade, C., & Terray, L. (2010). Twentieth century Sahel rainfall variability as simulated by the ARPEGE AGCM, and future changes. Climate Dynamics, 35, 75–94. Deser, C., & Blackmon, M. L. (1993). Surface climate variations over the North Atlantic Ocean during winter: 1900–1989. American Meteorological Society, 1743–1753. Eklundh, L., & Olsson, L. (2003). Vegetation index trends for the African Sahel 1982– 1999. Geophysical Research Letters, 30. Fensholt, R., & Rasmussen, K. (2011). Analysis of trends in the Sahelian ‘rain-use efficiency’ using GIMMS NDVI, RFE and GPCP rainfall data. Remote Sensing of Environment, 115, 438–451. Folland, C. K., Palmer, T. N., & Parker, D. E. (1986). Sahel rainfall and worldwide sea temperatures, 1901–85. Nature, 320, 602–607. Giannini, A., Saravanan, R., & Chang, P. (2003). Oceanic forcing of Sahel rainfall on interannual to interdecadal time scales. Science, 302, 1027–1030. Hagos, S. M., & Cook, K. H. (2008). Ocean warming and late-twientieth-century Sahel drought and recovery. Journal of Climate, 21, 3797–3814. Heinemann, B.Open University Course Team. (1989). Ocean circulation. Oxford, U.K. Pergamon Press. Herman, A., Kumar, V. B., Arkin, P. A., & Kousky, J. V. (1997). Objectively determined 10- day African rainfall estimates created for famine early warning systems. International Journal of Remote Sensing, 18, 2147–2159. Herrmann, S. M., Anyamba, A., & Tucker, C. J. (2005). Recent trends in vegetation dynamics in the African Sahel and their relationship to climate. Global Environ- mental Change-Human and Policy Dimensions, 15, 394–404. Heumann, B. W., Seaquist, J. W., Eklundh, L., & Jonsson, P. (2007). AVHRR derived phenological change in the Sahel and Soudan, Africa, 1982–2005. Remote Sensing of Environment, 108, 385–392. Hickler, T., Eklundh, L., Seaquist, J. W., Smith, B., Ardo, J., Olsson, L., Sykes, M. T., & Sjostrom, M. (2005). Precipitation controls Sahel greening trend. Geophysical Research Letters, 32. Holben, B. N. (1986). Characteristics of maximum-value composite images from temporal AVHRR data. International Journal of Remote Sensing, 7, 1417–1434. Huber, S., Fensholt, R., & Rasmussen, K. (2011). Water availability as the driver of vegetation dynamics in the African Sahel from 1982–2007. Global and Planetary Change, 76, 186–195. James, M. E., & Kalluri, S. N. V. (1994). The Pathfinder AVHRR land data set — An improved coarse resolution data set for terrestrial monitoring. International Journal of Remote Sensing, 15, 3347–3363. Janicot, S., Harzallah, A., Fontaine, B., & Moron, V. (1998). West African monsoon dynamics and eastern equatorial Atlantic and Pacific SST anomalies (1970–88). Journal of Climate, 11, 1874–1882. Jarlan, L., Tourre, Y. M., Mougin, E., Philippon, N., & Mazzega, P. (2005). Dominant patterns of AVHRR NDVI interannual variability over the Sahel and linkages with key climate signals (1982–2003). Geophysical Research Letters, 32. Jury, M. R., & Mpeta, E. J. (2009). African climate variability in the satellite era. Theoretical and Applied Climatology, 98, 279–291. Lamb, P. J. (1978). Large-scale tropical Atlantic surface circulation patterns associated with sub-Saharan weather anomalies. Tellus, 30, 240–251. Le Houérou, H. N. (1989). The grazing land ecosystems of the African Sahel. Berlin; New York: Springer-Verlag. Lebel, T., & Ali, A. (2009). Recent trends in the Central and Western Sahel rainfall regime (1990–2007). Journal of Hydrology, 375, 52–64. Los, S. O. (1998). Estimation of the ratio of sensor degradation between NOAA AVHRR channels 1 and 2 from monthly NDVI composites. IEEE Transactions on Geoscience and Remote Sensing, 36, 206–213. Lu, J. (2009). The dynamics of the Indian Ocean sea surface temperature forcing of Sahel drought. Climate Dynamics, 33, 445–460. Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M., & Francis, R. C. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society, 78, 1069–1079. Mohino, E., Janicot, S., & Bader, J. (2010). Sahel rainfall and decadal to multi-decadal sea surface temperature variability. Climate Dynamics, 1–22. Moron, V. (1994). Guinean and Sahelian rainfall anomaly indexes at annual and monthly scales (1933–1990). International Journal of Climatology, 14, 325–341. Nicholson, S. E., Davenport, M. L., & Malo, A. R. (1990). A comparison of the vegetation response to rainfall in the Sahel and East-Africa, using Normalized Difference Vegetation Index from NOAA AVHRR. Climatic Change, 17, 209–241. Nicholson, S. E., Some, B., & Kone, B. (2000). An analysis of recent rainfall conditions in West Africa, including the rainy seasons of the 1997 El Nino and the 1998 La Nina years. Journal of Climate, 13, 2628–2640. Oba, G., Post, E., & Stenseth, N. C. (2001). Sub-saharan desertification and productivity are linked to hemispheric climate variability. Global Change Biology, 7, 241–246. Palmer, T. N. (1986). Influence of the Atlantic, Pacific and Indian Oceans on Sahel rainfall. Nature, 322, 251–253. Philippon, N., Jarlan, L., Martiny, N., Camberlin, P., & Mougin, E. (2007). Characterization of the interannual and intraseasonal variability of West African vegetation between 1982 and 2002 by means of NOAA AVHRR NDVI data. Journal of Climate, 20, 1202–1218. Pinzon, J. E., Brown, M. E., & Tucker, C. J. (2005). Hilbert–Huang transform and its applications. In N. E. Huang (Ed.), Interdisciplinary mathematical sciences, v. 5, (pp. xii, 311 p.). Singapore; Hackensack, NJ; London: World Scientific. Plisnier, P. D., Serneels, S., & Lambin, E. F. (2000). Impact of ENSO on East African ecosystems: a multivariate analysis based on climate and remote sensing data. Global Ecology and Biogeography, 9, 481–497. Prince, S. D. (1991). A model of regional primary production for use with coarse resolution satellite data. International Journal of Remote Sensing, 12, 1313–1330. Prince, S. D. (1991). Satellite remote-sensing of primary production — Comparison of results for Sahelian grasslands 1981–1988. International Journal of Remote Sensing, 12, 1301–1311. Prince, S. D., De Colstoun, E. B., & Kravitz, L. L. (1998). Evidence from rain-use efficiencies does not indicate extensive Sahelian desertification. Global Change Biology, 4, 359–374. Propastin, P., Fotso, L., & Kappas, M. (2010). Assessment of vegetation vulnerability to ENSO warm events over Africa. International Journal of Applied Earth Observation and Geoinformation, 12S, 83–89. Raicich, F., Pinardi, N., & Navarra, A. (2003). Teleconnections between Indian monsoon and Sahel rainfall and the Mediterranean. International Journal of Climatology, 23, 173–186. Rasmussen, M. S. (1992). Assessment of millet yields and production in northern Burkina Faso using integrated NDVI from the AVHRR. International Journal of Remote Sensing, 13, 3431–3442. Reynolds, R. W., Rayner, N. A., Smith, T. M., Stokes, D. C., & Wang, W. Q. (2002). An improved in situ and satellite SST analysis for climate. Journal of Climate, 15, 1609–1625. Rowell, D. P. (2003). The impact of Mediterranean SSTs on the Sahelian rainfall season. Journal of Climate, 16, 849–862. Saji, N. H., Goswami, B. N., Vinayachandran, P. N., & Yamagata, T. (1999). A dipole mode in the tropical Indian Ocean. Nature, 401, 360–363. Seaquist, J. W., Hickler, T., Eklundh, L., Ardo, J., & Heumann, B. (2008). Disentangling the effects of climate and people on Sahel vegetation dynamics. Biogeosciences Discussions, 5, 3045–3067. Shanahan, T. M., Overpeck, J. T., Anchukaitis, K. J., Beck, J. W., Cole, J. E., Dettman, D. L., Peck, J. A., Scholz, C. A., & King, J. W. (2009). Atlantic forcing of persistent drought in West Africa. Science, 324, 377–380. Tucker, C. J., Pinzon, J. E., Brown, M. E., Slayback, D. A., Pak, E. W., Mahoney, R., Vermote, E. F., & El Saleous, N. (2005). An extended AVHRR 8-km NDVI dataset compatible with MODIS and SPOT vegetation NDVI data. International Journal of Remote Sensing, 26, 4485–4498. Tucker, C. J., Vanpraet, C. L., Sharman, M. J., & Vanittersum, G. (1985). Satellite remote- sensing of total herbaceous biomass production in the Senegalese Sahel — 1980– 1984. Remote Sensing of Environment, 17, 233–249. 3284 S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285
  • 10. UCAR (2011). http://www2.ucar.edu/news/backgrounders/arctic-oscillation-pineapple- express-weather-maker-glossary (accessed Jan 2011) Vermote, E., & Kaufman, Y. J. (1995). Absolute calibration of AVHRR visible and near- infrared channels using ocean and cloud views. International Journal of Remote Sensing, 16, 2317–2340. Wang, G. L. (2003). Reassessing the impact of North Atlantic Oscillation on the sub- Saharan vegetation productivity. Global Change Biology, 9, 493–499. Ward,M. N. (1998). Diagnosis and short-lead time prediction ofsummerrainfall intropical North Africa at interannual and multidecadal timescales. Journal of Climate, 11, 3167. Williams, C. A., & Hanan, N. P. (2011). ENSO and IOD teleconnections for African ecosystems: evidence of destructive interference between climate oscillations. Biogeosciences, 8, 27–40. Wolter, K., & Timlin, M. S. (1998). Measuring the strength of ENSO events: How does 1997/98 rank? Weather, 53, 315–324. 3285S. Huber, R. Fensholt / Remote Sensing of Environment 115 (2011) 3276–3285