Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

826 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

Derivation and Evaluation of Global 1-km Fractional Vegetation Cover Data for
Land Modeling
XUBIN ZENG, ROBERT E. DICKINSON, ALISON WALKER, AND MUHAMMAD SHAIKH
Institute of Atmospheric Physics, The University of Arizona, Tucson, Arizona

RUTH S. DEFRIES
Department of Geography, University of Maryland, College Park, College Park, Maryland

JIAGUO QI
Department of Geography, Michigan State University, East Lansing, Michigan

(Manuscript received 3 May 1999, in final form 23 July 1999)

ABSTRACT
Fractional vegetation cover (s y ) is needed in the modeling of the land–atmosphere exchanges of momentum,
energy, water, and trace gases. From global 1-km, 10-day composite Advanced Very High Resolution Radiometer
normalized difference vegetation index (NDVI) data from April 1992 to March 1993, global 1-km s y is derived
based on the annual maximum NDVI value for each pixel in comparison with the NDVI value that corresponds
to 100% vegetation cover for each International Geosphere–Biosphere Program land cover type. This dataset is
pixel dependent but season independent, with the seasonal variation of vegetation greenness in a pixel accounted
for by the leaf area index. The authors’ algorithm is found to be insensitive to the use of a specific land cover
classification. In comparison with an independent dataset derived by DeFries et al. by using a more sophisticated
statistical approach, the current dataset has a similar spatial distribution but systematically smaller s y (particularly
over shrublands and barren land cover). It also gives s y values that overall are consistent with those derived
from higher-resolution aircraft and satellite data over Arizona and field-survey data over Germany.

1. Introduction is a cleanly obtainable remote sensing product for frac-


tional vegetation that can be translated directly into us-
The land component of climate models represents
age in a climate model without implicitly changing its
many important processes that control the transfers of
definition in the process.
water and energy to the atmosphere. Because the ab-
sorption of solar energy is a dominant process, surface Most land models up to now have used the undoc-
properties (such as the presence of vegetation) that are umented and unvalidated estimates for fractional veg-
related to this absorption must be included in some de- etation included in the Biosphere–Atmosphere Transfer
tail. Hence, the fractional vegetation cover, as first in- Scheme (BATS) (Dickinson 1993). These estimates
troduced by Deardorff (1978), is an important element originated in the early 1980s and were derived by look-
of climate models. In the absence of satellite remote ing out the window. In addition to maximum values, a
sensing, its specification from observations has been temperature-based seasonal variation was employed in
problematical. With satellite observations, it should be global model implementations. The first satellite anal-
a relatively simple parameter to obtain. The variety of yses (Gutman and Ignatov 1998) thus have employed
detailed definitions of fractional vegetation employed the Advanced Very High Resolution Radiometer
by modeling and remote sensing, however, can lead to (AVHRR) normalized difference vegetation index
a mismatch between what was estimated from remote (NDVI) data to estimate seasonally varying fractional
sensing and what is used in a model. Developed here vegetation. Land models also require a leaf area index
(LAI) to determine surface energy fluxes over canopy-
covered surfaces. This LAI can affect how much solar
radiation is reflected and hence can affect NDVI. Al-
Corresponding author address: Xubin Zeng, Department of At-
mospheric Sciences, The University of Arizona, P.O. Box 210081,
though many climate models have included a seasonal
Tucson, AZ 85721-0081. variation of LAI, the limited information available from
E-mail: xubin@atmo.arizona.edu the NDVI product precludes deriving seasonal varia-

q 2000 American Meteorological Society


JUNE 2000 ZENG ET AL. 827

tions of both fractional vegetation and LAI. Hence, Gut- r2 2 r1


man and Ignatov (1998) prescribed a fixed value for N5 , (1)
r2 1 r1
LAI. Alternatively, Sellers et al. (1996) and other au-
thors assumed that spatial and seasonal variations of where r1 is the visible reflectance at 0.58–0.68 mm
NDVI all were given by variations of LAI. Any of the (AVHRR channel 1), and r 2 is the near-infrared reflec-
four possible assumptions (i.e., whether fractional veg- tance at 0.73–1.1 mm (channel 2). The fractional veg-
etation and/or LAI vary seasonally or are kept constant) etation cover s y can be linked to N in at least three
could be used in a climate model, but we argue that different ways.
fixing fractional vegetation and allowing LAI to vary Consistent with model practice in representing grid-
is a realistic assumption from a modeling viewpoint and box flux as the area-weighted average from vegetated
can be supported by current observational data.1 In this area and bare soil, a quantity measured by satellite data
context, fractional vegetation is viewed as being deter- (f ) for each pixel can be interpreted as the sum of
mined by types of vegetation and long-term edaphic and contributions from vegetated area (s y ) and bare soil (1
climatic controls, whereas LAI, as provided by model 2 s y ) (i.e., the ‘‘linear-mixing’’ assumption):
simulations (e.g., Dickinson et al. 1998) or remote sens-
ing (e.g., Myneni et al. 1997), includes all the season- f 5 f y s y 1 (1 2 s y )f s , (2)
ality of canopies, ranging from near-zero values in the where the subscripts y and s denote values over fully
more extreme cases of annual or harvested vegetation vegetated area and bare soil, respectively (in this and
to full-canopy values of 5–10. A small mismatch be- all other equations in the paper). Application of this
tween models and observations still may occur as a equation directly to N yields the simplest formulation
result of diurnally varying shading because a satellite for s y :
would measure a fraction of solar radiation reflected
from nonvegetated surfaces at the time of satellite over- N 2 Ns
sy 5 . (3)
pass, whereas a model definition of nonvegetated sur- Ny 2 Ns
faces would refer better to conditions averaged over the
Alternatively, (2) could be applied to r1 and r 2 sep-
day. Validation of fractional vegetation is problematical
arately because they are proportional to energy flux.
because of its requirement for information at the scale
of individual plant elements. DeFries et al. (1999) pre- Then (1) and (2) would give
viously have addressed this issue through use of the N 2 Ns
30-m Landsat Multispectral Scanner (MSS) data. sy 5 , (4)
Ny 2 Ns 1 (1 2 a)(N 2 Ny )
Because they generally occur on scales smaller than
climate model grids, the fractional vegetated or non- where a is equal to (r1 1 r 2)y /(r1 1 r 2)s . Equation (4)
vegetated surfaces generally will be used as part of the becomes the same as the linear (3) when a is equal to 1.
model parameterization of subgrid-scale heterogeneity. Another nonlinear (in N) formulation (Choudhury et
It can be introduced either in terms of complex surfaces al. 1994; Baret et al. 1995; Wittich 1997) is
as in BATS (Dickinson et al 1993) or in terms of a

1 2
Ny 2 N
b
mosaic of simpler surfaces (Avissar and Pielke 1989;
sy 5 1 2 , (5)
Koster and Suarez 1992). Either approach may be ap- Ny 2 Ns
propriate in a given situation, depending on the spatial
scale of the heterogeneity in comparison with that of which becomes (3) when the exponent b is equal to 1.
individual plants. When the required scaling information Although some studies (e.g., Myneni and Williams
is absent (as is generally the case), however, the mosaic 1994; Carlson and Ripley 1997) reported a nonlinear
approach may be preferred for climate models because relationship between N and s y , many others [e.g., Wit-
of its relatively easy implementation. tich (1997); Gutman and Ignatov (1998); and references
Section 2 discusses the derivation of global 1-km frac- therein] found the linear relationship (3) to be adequate.
tional vegetation cover, and section 3 evaluates this da- A more-elaborate mixture model (DeFries et al. 1999)
taset using higher-resolution aircraft and satellite data provides products as an area-weighted (linear) average
and field-survey data. for each of the 30 metrics (including N) and derives
continuous fields of vegetation characteristics. For a
2. Global fractional vegetation cover data specific area when surface data are available, a nonlinear
relationship (4) or (5) or an even more–sophisticated
a. Formulation approach (e.g., Gillies and Carlson 1995; Hall et al.
The normalized difference vegetation index (denoted 1997), might be more appropriate. For global applica-
as N hereinafter for brevity) is computed as tions, however, it may be difficult to justify anything
more elaborate than the linear relationship (3). Largely
because of a lack of data, it already is difficult to de-
1
Derivation of season-independent fractional vegetation cover also termine N y and N s in (3). In fact, a universal value of
was discussed by G. Gutman in an unpublished draft in 1995 (G. 0.52 is adopted for N y and 0.04 for N s in Gutman and
Gutman 1999, personal communication). Ignatov (1998). If (4) or (5) were used, it would be even
828 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

more difficult to determine the additional parameter a TABLE 1. Pixel numbers [relative to the total pixel number of about
1.3 3 10 8 over land (i.e., categories 1–14 and 16)], Nc, y used in (6),
or b. and the mean fractional vegetation cover for each IGBP land cover.

IGBP land cover Pixels (%) N c, y sy


b. Input data
1. Evergreen needleleaf forests 5.03 0.63 0.90
The global 1-km, 10-day composite AVHRR N data 2. Evergreen broadleaf forests 9.39 0.69 0.93
from April 1992 to March 1993 as obtained from the 3. Deciduous needleleaf forests 1.52 0.63 0.92
4. Deciduous broadleaf forests 2.50 0.70 0.90
Earth Resources Observation Systems (EROS) Data 5. Mixed forests 4.86 0.68 0.88
Center (EDC) archive (Eidenshink and Faundeen 1994) 6. Closed shrublands 2.01 0.60 0.72
are used. To compute 10-day maximum N composites, 7. Open shrublands 13.96 0.60 0.39
they first applied orbital stitching, radiometric calibra- 8. Woody savannas 7.87 0.62 0.86
tion, and geometric registration of all five AVHRR chan- 9. Savannas 7.21 0.58 0.81
10. Grasslands 8.53 0.49 0.71
nels. Then, the maximum N value for each pixel and 11. Permanent wetlands 1.02 0.56 0.85
the corresponding values for each channel were selected 12. Croplands 10.89 0.61 0.86
for a 10-day period (or three composites per month). 13. Urban and built-up lands 0.20 0.62 0.74
The retention of the highest N helps to reduce the num- 14. Cropland/natural vegetation 10.80 0.65 0.85
15. Snow and ice
ber of cloud-contaminated pixels. Last, atmospheric cor- 16. Barren 14.22 0.60 0.11
rections for ozone and Rayleigh scattering of the 17. Water bodies
AVHRR visible and near-infrared channels (channels 1
and 2) were applied, and the maximum N value was
recomputed. No water vapor or aerosol correction was
water pixels also are defined as water from the N data.
applied, however, because of a lack of consensus as to
Only 1.25% of snow pixels are classified as vegetated
a feasible method. Data masks include missing data over
area based on N (with N p,max $ 0.05). About 40 000
land, ocean pixels, Goode’s interrupted area, and solar
pixels (or about 0.03% of the land) are classified as
zenith angles greater than 808.
missing data in the land cover data but have N p,max great-
Evaluation of this dataset over the continental United
er than or equal to 0.05. To make the two datasets more
States by Zhu and Yang (1996) revealed that the max-
consistent, the IGBP land cover data are modified slight-
imum value composite process described above has a
ly as follows: nonvegetated pixels (i.e., in the IGBP
bias toward postnadir direction, particularly during the
categories of snow and ice, water, interrupted points, or
winter, and thus this bias may be mitigated to a certain
missing values) that have N p,max greater than or equal to
extent by extending the compositing period from 10
0.05 are redefined as the dominant vegetation type of
days to 1 month. One-month compositing also was
the adjacent nine pixels; land pixels (from the IGBP
found to reduce cloud contaminations of data effec-
land cover data) that have ocean flags based on the
tively.
NDVI data are redefined as ocean. This slightly mod-
Also used are the global 1-km land cover data (Love-
ified IGBP land cover dataset is used for subsequent
land and Belward 1997), as derived from the global
work.
1-km N data spanning 12-month period (April 1992–
March 1993) (i.e., the same N data as mentioned earlier).
Their mapping uses unsupervised classification with po- c. Results
stclassification refinement using ancillary data (includ-
ing digital elevation data, ecoregions interpretations, A fractional vegetation cover s y for each pixel that
and country- or regional-level vegetation and land cover is independent of season (as discussed in section 1) is
maps). Of the six different land cover classification obtained by applying the annual maximum N (N p,max )
schemes they provide, the International Geosphere–Bio- to the linear (3):
sphere Program (IGBP) land cover classification (Bel- Np,max 2 Ns
ward 1996) (see Table 1), which is adopted by the Na- sy 5 , (6)
Nc,y 2 Ns
tional Aeronautics and Space Administration Earth Ob-
serving System (NASA EOS) program, is used here. where N c,y is the N value for each IGBP category that
Some other classifications also will be used in section corresponds to 100% vegetation cover. The use of N p,max
3 for comparison. implies that s y represents the annual maximum green
To evaluate the above two input data sources, the vegetation fraction for a given pixel. This use (in con-
maximum N for each pixel during the 12 months (de- trast to 10-day or monthly N) minimizes as much as
noted as N p,max ) is computed. Because the IGBP land possible the effects of cloud contamination.
covers were derived primarily from the same N data as Rules for determining N c,y and N s require the histo-
were used in this study, they are expected to agree with gram of N p,max for each IGBP category to be computed.
each other. Indeed, 99.25% of land pixels based on the The fifth percentile for the category 16 (barren or sparse-
land cover data have N p,max greater than 0.05 (i.e., the ly vegetated) is taken as N s (and this value is 0.05).
N value for bare soil; see section 2c), and 98.89% of Most of the other categories have their smaller NDVI
JUNE 2000 ZENG ET AL. 829

their s y values. The use of the 75th percentile implies


that 25% of the pixels have 100% vegetation cover (Fig.
1b,c). Obviously, the percentile used for N c,y should be
higher for shrublands (e.g., the 90th percentile used in
this study) than for forests. It is clear from Fig. 1 that
N c,y based on the 75th percentile differs by only 20.02
and 0.03, respectively, if the 50th or 90th percentile is
used instead. In other words, N c,y is insensitive to the
exact percentile used for a given vegetation type. Figure
1 also shows that s y is greater than 0.95 for 50% of the
pixels, as expected for evergreen broadleaf forest, and
s y is less than 0.6 for only 1% of the pixels that are
located at the forest edge, are partially deforested, or
are misclassified.
Figure 2 shows the derived 1-km fractional vegetation
cover and the corresponding IGBP land cover and to-
pography data (from the global 30–arc second elevation
dataset, available at the time of writing at http://
edcwww.cr.usgs.gov/landdaac) over the southwestern
United States. The dominant land cover is shrubland,
FIG. 1. (a) Histogram, (b) cumulative frequency distribution, and followed by grassland and evergreen needleleaf forest
(c) fractional vegetation cover of N p,max for evergreen broadleaf forest (Fig. 2b). Los Angeles is the largest urban area in Fig.
pixels. 2b, and Tucson also can be found, at the lower-right
part of Fig. 2b. The fractional vegetation cover s y for
the forest is higher than that for the grassland, the s y
values in winter and thus have larger uncertainties of which, in turn, on average is higher than that for
caused by cloud contamination and atmospheric effects shrubland (Fig. 2a). In particular, the western half of
than in summer. Thus, it is not practical to determine the Baja California Peninsula facing the Pacific Ocean
N s separately for each category, so we assume N s 5 is covered by closed shrubland and has a much higher
0.05 for all the IGBP categories. s y than has the eastern half that is covered by open
To determine what percentiles should be used for N c,y , shrubland. Figure 2c shows that evergreen needleleaf
a commercial imagery database (http:// forest exists only at high altitudes along the Mogollon
www.terraserver.microsoft.com) of 1 terabyte of data rim over central Arizona, the Sierra Nevada over Cal-
from mostly over the continental United States and ifornia, and the mountains over Baja California. The
western Europe was extensively explored. This database Grand Canyon can be seen clearly as the green sinu-
combines the 1-m-resolution digitized aerial photo- soidal curve in the upper-middle part of Fig. 2c, along
graphs from the United States Geological Survey which s y is relatively small (Fig. 2a). These results are
(USGS) and the 2-m-resolution declassified images consistent with the authors’ experiences as local resi-
from sophisticated Russian mapping satellites. Frac- dents.
tional vegetation cover for each selected image was es- Figure 3 shows the global 10-km-averaged s y . Frac-
timated visually by three authors independently, and tional cover is nearly 100% over areas with extensive
their mean values were used. The dominant land cover forest cover (e.g., the Amazon basin) and is less than
types of these images were identified using the global 10% over deserts (e.g., the Sahara Desert). As expected,
1-km land cover data. Because these images were taken the s y values are relatively low over tundra (over North-
at times that usually do not correspond to the peak of ern Hemisphere high latitudes), the western United
the growing season, the vegetation cover estimates are States, western South America, western South Africa,
of limited accuracy. On this basis, N c,y is taken as the and in the interior of Australia. Figure 4 and Table 1
75th percentile for the IGBP categories 1–5, 8–12, and show the averaged s y for each IGBP land cover type.
14 and the 90th percentile for categories 6 (closed shrub- It is about 90% for forests, 39% for open shrubland,
land) and 13 (urban and builtup lands). Here N c,y for 11% for barren and sparsely vegetated areas, and about
category 7 (open shrubland) or 16 (barren or sparsely 80% for other vegetation types. Figure 5 shows the zon-
vegetated) is taken to be the same as that for category ally averaged s y (over land). Consistent with Fig. 3, the
6. The N c,y values so determined are given in Table 1. maximum s y occurs over the tropical broadleaf forests
The effect of uncertainties in N c,y and N s on fractional and over needleleaf forests at high latitudes. Lower s y
vegetation cover will be addressed in section 3. values over subtropical areas (Fig. 5) correspond to the
The above procedure is illustrated by Fig. 1, which barren or semiarid areas in Fig. 3. Lower s y values near
shows the histogram and cumulative frequency distri- 478S in Fig. 5 are consistent with those in southern Chile
bution of N p,max for evergreen broadleaf forest pixels and and Argentina in Fig. 3.
830 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

FIG. 2. (a) Derived 1-km fractional vegetation cover, (b) IGBP land cover types (categories are
listed in Table 1), and (c) 1-km topography over the southwestern United States.
JUNE 2000
ZENG ET AL.

FIG. 3. Global 10-km-averaged fractional vegetation cover based on the 1-km data.
831
832 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

FIG. 4. Averaged fractional vegetation cover for each IGBP land cover type (see Table 1). Results based on the IGBP, BATS, USGS, and
SiB2 classifications are denoted by the dark green, light green, blue, and red bars, respectively. Results from DeFries et al. (1999) are denoted
by the gold bars.

3. Evaluation of the dataset corresponding changes of s y are only 0.03 and 20.04,
respectively.
Past attempts to validate global datasets for fractional
The sensitivity of the derived values to the use of four
vegetation have been limited by the formidable diffi-
culties of such an effort. A comprehensive attempt at different land cover classification schemes is addressed
validation should recognize that the computation of frac- next. Global 1-km land cover data based on these four
tional vegetation cover s y from (6) is insensitive to N s schemes are all obtained from EDC (see section 2b).
because N s appears in both the numerator and denom- These schemes include the IGBP, BATS (Dickinson et
inator. For instance, for category 1 (evergreen needleleaf al. 1993), USGS (see http://edcwww.cr.usgs.gov/land-
forest), changing N s by 60.03 would change the mean daac/glcc/globepint.html), and the revised version of the
s y by only 60.01. In contrast, because N c,y appears only Simple Biosphere model (SiB2) (Sellers et al. 1996).
in the denominator, incorrect classification of pixels, They represent different aggregations of the same data
particularly between categories with significantly dif- used in the unsupervised land cover classification
ferent N c,y values (e.g., grasslands versus broadleaf for- (Loveland and Belward 1997). They all have forest
est), might introduce larger errors. For instance, an error types similar to those shown in Table 1 but differ in
of 0.04 (equal to the mean absolute difference of N c,y their inclusion or omission of other vegetation types
between category 1 and other categories) would change (particularly cropland–natural vegetation mosaic and
the mean s y by 60.07. As shown in Fig. 1, N c,y based tundras). Just as in using the IGBP scheme in section
on the 75th percentile differs by only 20.02 and 0.03, 2, N c,y is taken as the 75th percentile based on the his-
respectively, for evergreen broadleaf forest from its val- togram of N p,max for each category in the other three
ue if the 50th and 90th percentiles are used instead. The schemes with the following exceptions. For the BATS
JUNE 2000 ZENG ET AL. 833

FIG. 5. Zonally averaged fractional vegetation cover (over land). Results based on the IGBP,
USGS, BATS, and SiB2 classifications are denoted by the solid, dotted, dashed, and dot–dashed
lines, respectively. Results from DeFries et al. (1999) are denoted by the double-dotted–dashed
lines.

scheme, N c,y is taken as the 90th percentile for tundra Next, the dataset is compared with that independently
and evergreen and deciduous shrubs. The N c,y value derived by DeFries et al. (1999) using a more sophis-
(0.57) for evergreen shrub also is used for desert and ticated approach. First they derived linear discriminants
semidesert (categories 8 and 11). For the USGS scheme, for input into a linear mixture model from 30 metrics
N c,y is taken as the 90th percentile for urban, mixed that represent the annual phenological cycle. Then they
shrubland/grassland, and herbaceous, woody, and mixed determined coefficients (or endmember values) using
tundra. The N c,y value for mixed tundra is used for bare training data derived from a global network of Landsat
ground tundra, and N c,y for mixed shrubland/grassland MSS scenes. Last they applied a linear mixture model
is used for shrubland and barren. For the SiB2 scheme, to derive global continuous fields in terms of life form
N c,y is taken as the 90th percentile for dwarf trees and (percent woody vegetation, percent herbaceous vege-
shrubs, and this value also is used for shrubs with bare tation, and bare ground), leaf type, and leaf duration.
soil. Equation (6) then is used to compute fractional Here, the sum of percent woody and herbaceous veg-
vegetation cover for each pixel using each of these etations is compared with the fractional vegetation cover
schemes. from the current study. These two datasets give similar
Figure 4 shows that the four different classification variation of averaged s y from one IGBP land cover type
methods described above yield very similar results for to another (Fig. 4). They also give similar meridional
most of the vegetation categories. For shrublands and variation of zonally averaged s y (Fig. 5). To compute
barren (IGBP categories 6, 7, and 16), the USGS scheme their correlation, these data were averaged into 0.58 3
yields a higher s y than do the others, because N c,y is 0.58 bins, because DeFries et al. (1999) sampled data
just 0.47 (derived based on the 90th percentile for the every few pixels. The correlation coefficient is 0.85,
USGS mixed shrubland/grassland) versus 0.60 (derived again indicating a good agreement.
based on the 90th percentile for the IGBP closed shrub- The s y from DeFries et al. (1999) is systematically
land). For grassland (IGBP category 10), the SiB2 larger than that of this study. (Figs. 4 and 5). Their
scheme yields a lower s y than do the others, because it fractional covers for closed and open shrublands (IGBP
puts grassland and agriculture (or short vegetation) into types 6 and 7) are higher by 0.15 (Fig. 4). Over the
the same category and hence uses a higher N c,y (0.61) Sahara desert, 100% bare soil is assigned in DeFries et
for grassland than that used in the IGBP scheme (0.49; al. (1999) for pixels identified as barren (type 16) in the
see Table 1). Figure 5 also demonstrates that these four IGBP land cover dataset to overcome the problem of
schemes give very similar zonally averaged results. This unrealistically high s y derived for those pixels. Despite
analysis indicates that the derived fractional vegetation this, their s y for the barren type is still 0.05 higher than
cover is insensitive to the use of a specific land cover this study’s (Fig. 4). Because shrubs and barren ground
classification method because of the use of percentiles occupy about 30% of the total land surface (Table 1),
(instead of absolute values) to determine N c,y for each it is important to resolve these differences. Their s y for
category. the urban area (type 13) is much larger than this study’s
834 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

(Fig. 4), but its effect on modeling is small because this the monsoon season) show greener cover. Table 2 shows
category represents only 0.2% of the total land surface that the mean s value based on the 3-m data is twice
(Table 1). Comprehensive validation data, if available, as large as that using the 5-m data. Because the 1-m
could be used to reduce these differences by relaxing and 5-m data were obtained by the same instrument in
the assumption of full vegetation cover over their ho- the same day, they show the same values. The only
mogeneous Landsat MSS training scenes in DeFries et difference is that the 5-m data have a larger swath width.
al. (1999) and/or by decreasing the N c,y in (6). For this reason, results based on the 1-m data are not
The current study’s data could be compared also with shown in Fig. 6 or Table 2.
the maximum fractional vegetation cover derived by To evaluate the AVHRR data, the high-resolution s
Gutman and Ignatov (1998). Because they prescribe N c,y values described above are averaged into 1-km boxes
in (6) as a constant value (0.52) for all pixels, their and their mean, minimum, and maximum values then
fractional covers would be larger for all IGBP categories are given in Table 2. Because only 2 days of high-
except grassland, based on the N c,y values in Table 1 resolution data are available, it is impossible directly to
(assuming the same constant value is applied to the 1-km evaluate s y based on the annual maximum N in (6).
data). As mentioned earlier, another important differ- Instead, the N values for the same 10-day periods (but
ence is that they derive fractional cover as a function in 1992) are used in (6), and the resulting statistical
of pixel and season based on the assumption of a con- quantities are shown in Table 2 for comparison. Table
stant LAI, but the current study’s s y is independent of 2 shows that the current s is larger by 0.07 than that
season (and LAI variation is used to explain all the using 3-m or 5-m resolution data, on average. To un-
seasonal variation of vegetation greenness in a pixel). derstand this difference, Fig. 7 shows the monthly pre-
Next, high-resolution (1–5 m) data over southern Ar- cipitation observed at the Tucson Airport by the Na-
izona are used to evaluate the data. Twelve channels of tional Weather Service. Because the 1992 spring
data were obtained by the Landsat Thematic Mapper (March–May) is much wetter (87 mm of precipitation)
(TM) simulator aboard an aircraft on 29 May (yearday than the 1996 spring (only 8 mm), it is reasonable to
150) 1996. Flying over two different altitudes yields have a larger s, based on the 1992 N data in Table 2.
pixel resolutions of 1 and 5 m, respectively. Twelve A larger s based on the August 1992 AVHRR data in
channels of data of 3-m resolution were also obtained Table 2 can be explained similarly, based on the pre-
on 12 August (yearday 224) 1997. The center of the cipitation difference between the 1992 and 1996 sum-
mapping area is 318319N, 1108089W over southern Ar- mers. In August, the minimum s using AVHRR data is
izona. The data then were corrected geometrically and larger than that using the 3-m data, and the maximum
georeferenced. Because N is most sensitive to red re- s using AVHRR data is smaller. This result might be
flectance variations (e.g., Huete et al. 1997), data from caused partially by the year-to-year variation of N (as
channel 5 (0.63–0.69 mm, equivalent to TM band 3 and implied by the precipitation difference in Fig. 7) but
similar to AVHRR channel 1) are used directly to obtain also might indicate the nonlinearity between N and s.
Obviously these data cannot be used to evaluate data
r5 2 rs from DeFries et al. (1999), but Table 2 shows that their
s5 , (7)
ry 2 rs fractional vegetation cover is larger than the current one,
based on the annual maximum N. Note that their number
where r 5 is the channel-5 digital number, and r s and r y of pixels for averaging is smaller because approximately
are the corresponding channel-5 values for bare soil and only every sixth 1-km pixel was sampled in their dataset.
fully vegetated pixels, respectively. Here s is used to The Landsat TM data for four days (15 January, 20
separate from s y that uses N p,max in (6). March, 21 April, and 8 June 1997) also were obtained
Visual inspection of the data shows that only the 1-m for a 28 3 28 area centered at 109.48W, 31.48N over
data are able to distinguish individual shrubs; with a southern Arizona. First, digital numbers for individual
swath width of 0.85 km, however, the mapping area of TM bands were converted to radiances at the top of the
the 1-m data is too small to evaluate the 1-km AVHRR atmosphere, based on Moran et al. (1995). These ra-
data. Therefore, the 1-m data are used to determine r s diances then were converted to surface radiances, based
and r y in (7) as the average channel-5 values over bare on the atmospheric correction code for this region (pro-
soil and fully vegetated pixels (as identified by visual vided by Alfredo Huete at the University of Arizona).
inspection). The same values are used for the 5-m data Next, the channel-3 (visible; 0.63–0.69 mm) and chan-
obtained in the same day. For the 3-m data (in a different nel-4 (near infrared; 0.76–0.90 mm) surface radiances
day), r s and r y are reestimated by using the 1-m data to were used to compute N [based on (1)]. Last, N values
help to identify bare and fully vegetated areas. The s are used to obtain
values at the 3- and 5-m resolutions are shown in Fig.
6. The green belts represent trees and grasses along the N 2 Ns
San Pedro River. In comparison with the 5-m data (ob- s5 . (8)
Ny 2 Ns
tained before the monsoon season in Arizona; see the
dashed line in Fig. 7), the 3-m data (obtained during Just as in (7), s is used to separate from s y that uses
JUNE 2000
ZENG ET AL.

FIG. 6. The fractional vegetation cover [defined based on (7)] using aircraft data at 3-m and 5-m resolutions over southern Arizona. The swath widths of the 3-m and 5-m data are 2.5 km
and 4.5 km, respectively. Only parts of the scenes are shown, for clarity.
835
836 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

is higher than that from the TM data. The large differ-


ence over grasslands is expected, because the greenness
of grasslands is known to be very sensitive to precip-
itation over a semiarid region (e.g., Arizona), and pre-
cipitation in January and March 1993 and April and
June 1992 (when the AVHRR data were obtained) and
preceding months was much higher than in the corre-
sponding months in 1997 (when the TM data were ob-
tained). The larger s values using the TM data over
forests and savannas, however, might be caused by the
use of the same N y for the TM data. Because grassland
is the dominant vegetation type for the aircraft scene in
Fig. 6, N y (of 0.4) is tuned primarily to grassland. If
different and higher (as expected from Table 1) N y val-
FIG. 7. Monthly precipitation at the Tucson Airport. The years of ues were used for forests and savannas, the difference
1992, 1993, 1996, and 1997 are denoted by the solid, dotted, dashed, of s between the TM and AVHRR data would be small-
and dot–dashed lines, respectively. The 50-yr (1949–98) average er. Similarly, if a different (and higher) N y is used for
monthly precipitation is denoted by the plus signs. shrublands, the difference of s between the TM and
AVHRR data would be as large as that for grasslands,
N p,max in (6). Here, N s can be obtained easily as about because both grasslands and shrublands are affected
0.05 by averaging N over bare soil (as identified by the strongly by precipitation. Table 3 also shows that the
1-m data). The N y is estimated roughly as 0.4 by forcing fractional cover from DeFries et al. (1999) is much high-
s on 8 June 1997 to be similar to that from the 5-m er than the current cover (based on N p,max ) for each
data on 29 May 1996. Its dependence on vegetation vegetation type. With an area average of 0.87, DeFries
category is not considered because land cover data for et al. probably overestimated fractional cover for this
these TM scenes are not available, nor are higher-res- semiarid region.
olution data at the same time (e.g., same 10-day period Last, the current data were compared with those over
and same year). five areas of 9–35 km 2 in western Germany (Wittich
To compare the TM and AVHRR data, the s values and Hansing 1995). These five sites are Wesermarsch
at 30-m resolution from (8) are mapped into the AVHRR (088249E, 538169N), covered by grassland; Altes Land
1-km boxes, and the maximum s values during the 4 (098459E, 538309N), used for fruit growing; Hildesh-
days for each 1-km box are shown in Fig. 8. The cor- eimer Börde (108079E, 528149N), used for arable farm-
responding s values were computed similarly by ap- ing; Rheinhessen (088129E, 498459N), used for vine-
plying the maximum AVHRR N values for the same growing; and Pfälzer Wald (078509E, 498199N), covered
four 10-day periods (but in 1992 and 1993) to (6). These with forest stands. The typical land cover portions for
values and the 1-km IGBP land cover data are shown these areas were estimated by Wittich and Hansing
also in Fig. 8. The mapping area in Fig. 8 represents a (1995) from agricultural, forestry, and statistical agen-
small portion of Fig. 2, and Tucson can be found at the cies. Table 4 compares these values with fractional veg-
upper-left portion of Fig. 8c. Table 3 gives the statistical etation covers averaged over pixels within 60.058
quantities from Fig. 8 along with those from s y that around the above points, based on the current data and
uses N p,max in (6) and from DeFries et al. (1999). Both those of DeFries et al. (1999). Because approximately
Fig. 8 and Table 3 show that the s values from the only every sixth pixel was sampled in DeFries et al.
AVHRR data are larger than the TM data over grass- (1999), their number of pixels for averaging is much
lands and shrublands but smaller over forests and sa- smaller than that of this study (Table 4). Table 4 shows
vannas. Because grasslands and shrublands cover 73% that the current values are within 0.07 of estimates from
of the area (Table 3), the mean s from the AVHRR data Wittich and Hansing (1995) over four of the five areas,

TABLE 2. Comparison of fractional vegetation covers from the 5-m aircraft data (on 31 May 1996) and 3-m data (on 12 Aug 1997) vs
those based on the 10-day composite 1-km AVHRR N data (20-31 May and 10–19 Aug 1992, respectively) over southern Arizona. Results
using annual maximum N and from DeFries et al. (1999) also are shown. Minimum and maximum values from the 3-m and 5-m data are
obtained after averaging values into 1-km boxes.

5-m data AVHRR 3-m data AVHRR sy using


(May ’96) (May ’92) (Aug ’97) (Aug ’92) N p, max DeFries
Mean 0.17 0.24 0.35 0.42 0.65 0.88
Min 0.00 0.00 0.00 0.16 0.45 0.76
Max 0.54 0.57 0.82 0.59 0.84 1.0
Area (km 2 ) 108 114 77 114 114 15
JUNE 2000
ZENG ET AL.

FIG. 8. (a) 1-km-averaged fractional vegetation cover [defined based on (8)] using 4 days of 30-m Landsat TM data for a 28 3 28 area centered at 109.48W, 31.48N over southern Arizona,
(b) fractional vegetation cover using the maximum AVHRR N values for the same four 10-day periods (but in different years), and (c) IGBP land cover.
837
838 JOURNAL OF APPLIED METEOROLOGY VOLUME 39

TABLE 3. Mean fractional vegetation cover using data in Fig. 8. TABLE 4. Comparison of fractional vegetation covers over five areas
Results using annual maximum N and from DeFries et al. (1999) also in western Germany from Wittich and Hansing (1995; denoted as
are shown. WH) with those of the current study (‘‘ours’’) and DeFries et al.
(1999).
TM data AVHRR sy using
Land cover Area (%) (1997) (1992–93) N p, max DeFries Mean Min Max No.
Grasslands 41 0.37 0.56 0.67 0.91 Wesermarsch
Shrublands 32 0.28 0.35 0.39 0.72 Ours 1.00 1.00 1.00 73
Savannas 16 0.68 0.54 0.67 0.98 DeFries 1.00 1.00 1.00 2
Forests 10 0.87 0.69 0.79 0.99 WH 0.96
Whole area 100 0.44 0.50 0.59 0.87 Altes Land
Ours 0.91 0.75 1.00 72
DeFries 1.00 1.00 1.00 4
WH 0.98
and values from DeFries et al. (1999) are within 0.09 Hildesheimer
of estimates from Wittich and Hansing (1995) for all Ours 0.97 0.83 1.00 74
DeFries 0.90 0.79 1.00 2
five areas. The relatively large difference over Hildesh- WH 0.81
eimer Börde between the estimation of Wittich and Han- Rheinhessen
sing (1995) and this study’s [or that from DeFries et al. Ours 0.87 0.73 1.00 76
(1999)] could be caused by the year-to-year variation DeFries 0.94 0.85 1.00 3
of arable farming activities. Difference of fractional WH 0.86
covers averaged over all five areas is only 0.04 (or 0.05) Pfälzer
between Wittich and Hansing (1995) versus the current Ours 1.00 0.95 1.00 84
DeFries 0.97 0.94 1.00 2
study (or DeFries et al. 1999). WH 0.95

4. Conclusions
Fractional vegetation cover and LAI can be derived with estimates from 1-, 3-, and 5-m aircraft data over
from N data by specifying an explicit relationship be- southern Arizona and field-survey data over Germany.
tween them or by specifying one parameter (possibly Comparison with the 30-m Landsat TM data shows the
as a function of space). For use of 1-km data in land expected strong dependence on precipitation of vege-
modeling (e.g., through the mosaic approach), it is de- tation greenness over grasslands and shrublands in a
sirable to derive a s y that is pixel dependent but season semiarid region (e.g., Arizona).
independent, with the seasonal variation of vegetation None of the above evaluations entirely validates this
greenness in a pixel accounted for by time-varying LAI. dataset. The 1- to 5-m aircraft data might be relatively
The alternative of assuming fixed LAI and seasonally reliable but are not in the same year and, perhaps, more
varying fractional vegetation also may be useful, how- important, are available for a few days rather than for
ever. In particular, the seasonally variable vegetation an annual cycle. The 30-m Landsat TM data have a
fraction from Gutman and Ignatov (1998) has been dem- higher spatial resolution and might give better estimates
onstrated to improve results using the Mesoscale Eta of N than do the 1-km AVHRR data, but they have a
Model (Betts et al. 1997). similar uncertainty (as that of the AVHRR data) in de-
Using global 1-km, 10-day composite AVHRR N data riving fractional vegetation cover. The uncertainty of
from April 1992 to March 1993, a global 1-km s y was field-survey data in Germany is unknown, and the in-
derived, based on the annual maximum N value for each terannual variability of agricultural activities is not con-
pixel in comparison with the N value that corresponds sidered. To validate rigorously the current dataset or that
to 100% vegetation cover for each IGBP land cover of DeFries et al. (1999), high-resolution (e.g., 1–5 m)
type. This product has been evaluated by studying its aircraft data over different land cover types for a whole
sensitivity to land cover classification methods, by com- annual cycle would have to be obtained. Such data might
paring it with an independently derived dataset, and by be available from some of the NASA EOS validation
using higher-resolution data. Primarily because N c,y in sites in the future.
(6) (i.e., the N value that corresponds to 100% vege- Although more complete validation would be pref-
tation cover for each land cover category) is determined erable, this paper still represents the first effort to eval-
through the histogram of N values in each category, the uate carefully the derived dataset. If both fractional veg-
derived s y values are insensitive to the use of a specific etation cover and LAI are derived from N, uncertainty
land cover classification. When compared with that in- in s y can be partially compensated for in land modeling
dependently derived by DeFries et al. (1999) using a by the adjustment of LAI. For instance, given the same
more sophisticated (but still linear) mixture model, the LAI (with respect to pixel), a positive bias in s y would
current dataset gives a similar spatial distribution of s y decrease LAI with respect to vegetated area (i.e., LAI/
but consistently smaller s y values (particularly over s y ). Work currently is under way to address the differ-
shrublands and barren land cover). It also is consistent ence between satellite-derived s y data and those as-
JUNE 2000 ZENG ET AL. 839

sumed by climate models, and to address how this dif- S. T. Daughtry, 1994: Relations between evaporation coefficients
and vegetation indices studied by model simulations. Remote
ference and satellite data uncertainties may affect cli- Sens. Environ., 50, 1–17.
mate modeling. Deardorff, J. W., 1978: Efficient prediction of ground temperature
So far only one yr of AVHRR data has been analyzed, and moisture with inclusion of a layer of vegetation. J. Geophys.
in part because of data availability and the difficulty in Res., 83, 1889–1903.
handling over 150 Gb of data. Applying the same meth- DeFries, R. S., J. R. G. Townshend, and M. C. Hansen, 1999: Con-
tinuous fields of vegetation characteristics at the global scale. J.
od of DeFries et al. (1999) to global 8-km AVHRR data Geophys. Res., 104, 16 911–16 923.
from 1982 to 1993, DeFries et al. (2000) have dem- , M. C. Hansen, and J. R. G. Townshend, 2000: Global contin-
onstrated that the standard deviation of s y over these uous fields of vegetation characteristics: A linear mixture model
12 yr is less than 0.06 for each vegetation type used in applied to multi-year 8 km AVHRR data. Int. J. Remote Sens.,
21, 1389–1414.
their study. Because of the use of histograms rather than Dickinson, R. E., A. Henderson-Sellers, and P. J. Kennedy, 1993:
the absolute N values, the current approach is expected Biosphere–Atmosphere Transfer Scheme (BATS) version 1e as
to give a slightly smaller standard deviation. Work cur- coupled to the NCAR Community Climate Model. NCAR Tech.
rently is under way to quantify the year-to-year vari- Note NCAR/TN-3871STR, 72 pp. [Available from NCAR, P.O.
ability of the s y , based on multiyear global 1-km Box 3000, Boulder, CO 80307-3000.]
, M. Shaikh, R. Bryant, and L. Graumlich, 1998: Interactive
AVHRR data. canopies in a climate model. J. Climate, 11, 2823–2836.
Eidenshink, J., and J. Faundeen, 1994: The 1-km AVHRR global land
Acknowledgments. This work was supported by data set: First stages in implementation. Int. J. Remote Sens.,
NASA through its EOS IDS Program (429-81-22; 428- 15, 3443–3462.
Gillies, R. R., and T. N. Carlson, 1995: Thermal remote sensing of
81-22) and the Department of Energy under Grant DE- surface soil water content with partial vegetation cover for in-
FG02-91ER61216 (to The University of Arizona), by corporation into climate models. J. Appl. Meteor., 34, 745–756.
the VEGETATION Project (Cooperative Agreement 58- Gutman, G., and A. Ignatov, 1998: The derivation of the green veg-
5344-6-F806, 95/CNES/0403) at the USDA-ARS, etation fraction from NOAA/AVHRR data for use in numerical
through the NASA Landsat Science Team (Grant weather prediction models. Int. J. Remote Sens., 19, 1533–1543.
Hall, F. G., D. E. Knapp, and K. F. Huemmrich, 1997: Physically
S-41396-F to J. Qi), and by NASA Grant NAG56004 based classification and satellite mapping of biophysical char-
to R. S. DeFries. All input data are distributed by the acteristics in the southern boreal forest. J. Geophys. Res., 102,
Distributed Active Archive Center at the EROS Data 29 567–29 580.
Center. Dr. Alfredo Huete is thanked for his group’s Huete, A. R., H. Q. Liu, K. Batchily, and W. van Leeuwen, 1997: A
comparison of vegetation indices over a global set of TM images
assistance in using their atmospheric correction code for EOS-MODIS. Remote Sens. Environ., 59, 440–451.
and for helpful discussions. Drs. Randy Koster and Gar- Koster, R. D., and M. J. Suarez, 1992: A comparative analysis of two
ik Gutman and an anonymous reviewer are thanked for land surface heterogeneity representations. J. Climate, 5, 1379–
constructive and helpful comments. 1390.
Loveland, T. R., and A. S. Belward, 1997: The IGBP-DIS global
1-km land cover data set, DISCover: First results. Int. J. Remote
REFERENCES Sens., 18, 3289–3295.
Moran, M. S., R. D. Jackson, T. T. Clarke, J. Qi, F. Cabot, K. J.
Avissar, R., and R. A. Pielke, 1989: A parameterization of hetero- Thome, and B. L. Markham, 1995: Reflectance factor retrieval
geneous land surfaces for atmospheric numerical models and its from Landsat TM and SPOT HRV data for bright and dark
impact on regional meteorology. Mon. Wea. Rev., 117, 2113– targets. Remote Sens. Environ., 52, 218–230.
2136. Myneni, R. B., and D. L. Williams, 1994: On the relationship between
Baret, F., J. G. P. W. Clevers, and M. D. Steven, 1995: The robustness FPAR and NDVI. Remote Sens. Environ., 49, 200–211.
of canopy gap fraction estimates from red and near-infrared re- , R. R. Nemani, and S. W. Running, 1997: Estimation of global
flectances: A comparison of approaches. Remote Sens. Environ., leaf area index and absorbed PAR using radiative transfer mod-
54, 141–151. els. IEEE Trans. Geosci. Remote Sens., 35, 1380–1393.
Belward, A. S., 1996: The IGBP-DIS global 1 km land cover data Sellers, P. J., S. O. Los, C. J. Tucker, C. O. Justice, D. A. Dazlich,
set (DISCover): Proposal and implementation plans. IGBP-DIS G. J. Collatz, and D. A. Randall, 1996: A revised land surface
Working Paper No. 13, 61 pp. [Available from IGBP-DIS Office, parameterization (SiB2) for atmospheric GCMs. Part II: The
42 Ave. Gustave Coriolis 31057 Toulouse Cedex, France.] generation of global fields of terrestrial biophysical parameters
Betts, A. K., F. Chen, K. E. Mitchell, and Z. I. Janjic, 1997: As- from satellite data. J. Climate, 9, 706–737.
sessment of the land surface and boundary layer models in two Wittich, K.-P., 1997: Some simple relationships between land-surface
operational versions of the NCEP Eta Model using FIFE data. emissivity, greenness and the plant cover fraction for use in
Mon. Wea. Rev., 125, 2896–2916. satellite remote sensing. Int. J. Biometeor., 41, 58–64.
Carlson, T. N., and D. A. Ripley, 1997: On the relation between NDVI, , and O. Hansing, 1995: Area-averaged vegetation cover fraction
fractional vegetation cover, and leaf area index. Remote Sens. estimated from satellite data. Int. J. Biometeor., 38, 209–215.
Environ., 62, 241–252. Zhu, Z.-L., and L. Yang, 1996: Characteristics of the 1-km AVHRR
Choudhury, B. J., N. U. Ahmed, S. B. Idso, R. J. Reginato, and C. data set for North America. Int. J. Remote Sens., 17, 1915–1924.

You might also like