Yuan-LUE Models Site Comparison-20130423

advertisement
1
Target Journal: Remote Sensing of Environment
2
Global Comparison of Light Use Efficiency Models for Vegetation Gross Primary
3
Production based on Eddy Covariance Towers Data
4
Wenping Yuan1, Wenwen Cai1, Jiquan Chen2, Shuguang Liu3,4, Wenjie Dong1, Dan Liu1,
5
Jiangzhou Xia1, Yang Chen1, other coauthors
6
1State
7
Change and Earth System Science, Beijing Normal University, Beijing 100875, China;
8
2Department
9
3United
Key Laboratory of Earth Surface Processes and Resource Ecology, College of Global
of Environmental Sciences, University of Toledo, Toledo, OH 43606, USA;
States Geological Survey, Earth Resources Observation and Science Center, Sioux Falls,
10
South Dakota 57198, USA;
11
4
12
South University of Forestry and Technology, Changsha, Hunan 410004, China;
13
Abstract
State Engineering Laboratory of Southern Forestry Applied Ecology and Technology, Central
14
Predicting the gross primary productivity (GPP) of terrestrial ecosystems has been a
15
major challenge in quantifying the global carbon cycle. Among all the predictive methods, the
16
light use efficiency (LUE) model may have the most potential to adequately address the spatial
17
and temporal dynamics of GPP because of its theoretical basis and practicality. Many different
18
LUE models have been developed recently, but our understanding of the relative merits of
19
different models is poor. Using carbon flux measurements data from 155 eddy covariance sites, we
20
assessed the ability of seven LUE models and compared the major model algorithms.
21
Comparisons between modeled and observed GPP showed that the model performance
1
22
substantially differed among ecosystem types. In generally, high model performance was found
23
over the deciduous broadleaf forests and mixed forests, and low performance was observed over
24
the evergreen broadleaf forests and shrublands. Except CFlux model, over the cloudy and overcast
25
days, other six models showed a significant underestimation due to ignoring the impacts of diffuse
26
radiation on light use efficiency. Among seven models, CFlux and EC-LUE showed the better
27
performance than averaged level at the 76% and 75% sites. All models were examined for
28
simulating the interannual variability of GPP observations, and the higher simulation accuracy was
29
found at CFlux and EC-LUE models. Paired comparisons showed the models differences majorly
30
resulted from the differences of environmental regulations equations compared with the fraction of
31
PAR absorbed by the vegetation canopy, and especially water stress equations substantially
32
differed among seven models.
33
Key words
34
Gross primary production; Light use efficiency; CASA; C-Fix; CFlux; EC-LUE;
35
MODIS; VPM; VPRM;
36
37
38
1. Introduction
Terrestrial gross primary productivity (GPP) is the largest component flux of the global
39
carbon cycle, and is about 20 times greater than the amount of carbon from anthropogenic sources
40
(Canadell et al., 2007). Thus, even small fluctuations in GPP can cause large changes in the
41
airborne fraction of anthropogenic carbon and influence future climate warming scenarios
42
(Raupach et al., 2008). Terrestrial GPP also provides important societal services through provision
2
43
of food, fiber and energy. Regular monitoring of terrestrial GPP is therefore required to understand
44
and assess dynamics in the global carbon cycle, forecast future climate, and ensure long term
45
security in services provided by terrestrial ecosystems (Bunn and Goetz, 2006; Schimel, 2007).
46
Numerous of ecosystem models have been widely developed as a means of quantifying
47
spatial-temporal variations in GPP at large scales. However, different ecosystem models are
48
inconclusive regarding the magnitude and spatial distribution of GPP at the regional and global
49
scales. Recently, the model comparison, using standardized data from the North American Carbon
50
Program, showed none of the models consistently reproduce observed interannual variability
51
within measurement uncertainty because models can not represent the variability in spring
52
phenology, soil thaw and snowpack melting, and lagged response of ecosystems to extreme
53
climatic events (Keenan et al., 2012). Another study evaluated simulated daily average GPP from
54
26 models against estimated GPP at 39 eddy covariance flux tower sites across the United States
55
and Canada, and showed none of the models in this study match estimated GPP within the range
56
of uncertainty of observed fluxes, indicating the poor model performance (Schaefer et al., 2012).
57
These conclusions were supported by the previous comparison of 16 dynamic global vegetation
58
models that suggested the lowest estimation of global NPP (39.9 Pg C) by the Hybrid model was
59
approximately 50% smaller compared to what was estimated by the TURC model (80.5 Pg C)
60
(Cramer et al., 19999).
61
The light use efficiency (LUE) model based on the satellite data may have the most
62
potential to adequately address the spatial and temporal dynamics of GPP because of its
63
theoretical basis and practicality (Running et al., 2000). Independently and as a part of integrated
64
ecosystem models, the LUE approach has been used to estimate GPP and net primary production
3
65
(NPP) at various spatial and temporal scales (Potter et al., 1993; Prince and Goward, 1995;
66
Landsberg and Waring, 1997; Running et al., 2000; Xiao et al., 2004; Coops et al., 2005).
67
Numerous of studies have validated LUE models at regional and global scales in a variety of
68
major ecosystem types (Potter et al., 1993; Turner et al., 2006).
69
LUE models are often developed based on the unique assumptions driving by different
70
environmental variables, and formulate the processes controlling vegetation production in
71
different ways. Thus, there is diversity in both the complexity of the LUE model structure and
72
formulation though all of them follow the light use efficiency principle. Each model, therefore, is
73
a complex combination of scientific hypotheses and choices, and their estimates depend on these
74
inherent assumptions (Beer et al., 2010). Recent studies showed the large model uncertainties
75
within LUE models. For example, using satellite-based models, estimated GPP for North America
76
vary considerably between 12.2 and 18.7 Pg C yr-1 (Huntzinger et al., 2012). Available individual
77
model validations, however, are not sufficient to identify the sources of model differences and
78
shortcomings due to differences at validation datasets and driving variables. Therefore, in order to
79
move towards more robust estimates of vegetation production dynamics, it is necessary to first
80
compare estimates from a variety of LUE models, as well as evaluate estimates against consistent
81
and extensive measurements that are available (Running et al., 2004; Heinsch et al., 2006).
82
In this study, we evaluated how well seven different satellite-based models capture
83
spatio-temporal variations of GPP. The overarching goals of this study are to (1) examine the
84
model performance at the numerous of eddy covariance sites and (2) compare the temperature and
85
water response curves among the seven LUE models.
4
86
2. Model and data
87
2.1 Light use efficiency model
88
The LUE model is built on two fundamental assumptions (Running et al., 2004): (1) the
89
ecosystem GPP is directly related to absorbed photosynthetically active radiation (APAR) through
90
LUE, where LUE is defined as the amount of carbon produced per unit of APAR, and (2) realised
91
LUE may be reduced below its theoretical potential value by environmental stresses, such as low
92
temperatures or water shortages (Landsberg et al., 1986). The general form of the LUE model is:
93
GPP  PAR  fPAR  LUEmax  f (Ts ,Ws , )
(1)
94
where PAR is the incident photosynthetically active radiation (MJ m-2) per time period (e.g., day or
95
month), fPAR is the fraction of PAR absorbed by the vegetation canopy, LUEmax is the potential
96
LUE (g C m-2 MJ-1 APAR) without environment stress, f is a scalar varying from 0 to 1 that
97
represents the reduction of potential LUE under limiting environmental conditions, Ts and Ws are
98
temperature and water downward regulation scalars, and the multiplication of LUEmax and f is
99
realised LUE.
100
In this study, seven LUE models were selected to conduct the global comparison of
101
model performance, including CASA (Potter et al., 1993), CFix (Veroustraete et al., 2002), CFlux
102
(Turner et al., 2006; King et al., 2011), EC-LUE (Yuan et al., 2007, 2010), MODIS-GPP (Running
103
et al., 2000), VPM (Xiao et al., 2004) and VPRM model (Mahadevan et al., 2008). The detailed
104
model introduction and model operation can be found at Supplemental Online Material (SOM).
105
2.2 Data and method
106
LaThuile FLUXNET dataset was used in this study (http://www.fluxdata.org). Totally,
5
107
155 eddy covariance (EC) towers were included in this study, from six major terrestrial biomes:
108
evergreen broadleaf forest (EBF), deciduous broadleaf forest (DBF), mixed forest (MIF),
109
evergreen needleleaf forest (ENF), shrubland (SHR) and grassland (GRA) (Table S1; Figure S1).
110
Detailed information on data processing and site information (i.e. vegetation, climate and soils)
111
are available at the LaThuile FLUXNET Internet sites.
112
We examined the model performance using calibrated parameter values. Eighty percent
113
sites were selected to calibrate model parameters for each vegetation type, and other 20% sites
114
were used to validate models. This parameterization was repeated by 100 times, and the calibrated
115
parameter values were collected within Table 1. Annual simulated and observed GPP were
116
calculated in order to investigate the model performance on interannual variability of GPP. If
117
missing daily data was 20% of entire year data, the value of this year was indicated as missing.
118
For a site to be included for evaluating interannual variability, it had to have minimum of 3 years
119
of GPP observations and simulations. Based on this criterion, 100 sites consisting of 462 years
120
were included into the analysis (Table S1). We calculated the standard deviations of annual
121
averaged GPP observations and simulations for each site, and examined the correlations through
122
all 100 sites. Moreover, cloudiness index (CL), which was calculated by the ratio of PAR and
123
potential PAR, was used to indicate the friction of cloud cover. The days when CL is less than 0.3
124
were indicated as clear days, the CL ranges 0.3 to 0.6 for cloudy days and more than 0.6 were
125
indicated as overcast days. Similarly, water stress was separated into three levels (i.e. drought,
126
normal and wet conditions) based on the water stress scalars of seven models as the following
127
equations
6
Drought, Ws < Wsmin +
128
Normal , (Wsmin +
{ Wet
Wsmax −Wsmin
3
Wsmax −Wsmin
, Ws > Wsmin +
) < Ws <
3
2×(Wsmax −Wsmin )
(Wsmin +
2×(Wsmax −Wsmin )
3
(2)
3
129
where Wsmin and Wsmax are the minimum and maximum values of Ws at each site.
130
<<Table 1>>
131
)
Two pairwise comparisons were conducted on the model components in order to
132
investigate the differences of model structure. First, we identified the impacts of fraction of PAR
133
absorbed by the vegetation canopy on GPP simulations by comparing the two correlations:
134
(a) the correlation of simulated GPP among seven models;
135
(b) the correlation of potential light energy use (PLUE) (i.e. PAR×fPAR×LUEmax);
136
Then, we diagnosed the primary environment variables by the second pairwise
137
138
139
140
comparison:
(a) the correlation of realized light energy use only considering temperature stress
(RLUEtem) (i.e. PAR×fPAR×LUEmax×Ts);
(b) the correlation of realized light energy use only considering water stress (RLUEwater)
141
(i.e. PAR×fPAR×LUEmax×Ws).
142
2.3 Statistical analysis
143
The nonlinear regression procedure (Proc NLIN) in the Statistical Analysis System
144
(SAS, SAS Institute Inc., Cary, NC, USA) was applied to optimize the model parameters of seven
145
LUE models across the calibration sites. Four metrics were used to evaluate the performance of
146
the LUE models in this study, including coefficient of determination (R2), root mean square error
147
(RMSE), mean predictive error (PE, difference between mean observations and simulations), and
7
148
relative predictive error (RPE, the ratio between PE and mean observations).
149
3. Results and discussion
150
3.1 Comparison of model performance
151
All of seven LUE models showed the substantial difference of model performance
152
among various ecosystem types according to the R2, RMSE, PE and RPE (Figure S1). Over the
153
shrublands and evergreen broadleaf forests, almost all models showed obvious low performance
154
with low R2 and high RMSE. The highest model performance was observed over deciduous
155
broadleaf forests within seven LUE models. Parameter calibration significantly improved the
156
model performance over almost all ecosystem types (Figure 1; Table 1). Similarly, all of seven
157
calibrated models showed the highest model performance over the deciduous broadleaf forests,
158
intermediate over the evergreen needleleaf forests, mixed forests and grasslands, lowest at
159
shrublands and evergreen broadleaf forests (Figure 1). For a given vegetation type, models
160
performance differed among seven models. For example, CFlux and EC-LUE models were found
161
higher R2 and lower RMSE compared with other five models (Figure 1).
162
<<Figure 1>>
163
From the spatial scales, EC-LUE and CFlux models explained higher GPP variations
164
with determination coefficient of 0.55 and 0.44 respectively (Figure 2). All models appeared the
165
overestimation at the low GPP regions, and underestimation at the high GPP regions (Figure 2).
166
Moreover, we calculated the mean R2 and RMSE values of seven models at each site, and
167
compared the R2 and RMSE of individual model with means of seven models. On average, at 80%
168
and 75% sites, EC-LUE and CFlux models showed higher R2 compared with mean level of seven
8
169
models, while they showed the lower RMSE at 76% and 75% sites respectively (Figure 3).
170
<<Figure 2>>
171
<<Figure 3>>
172
Seven LUE models, expect CFlux model, significantly underestimated GPP at the
173
overcastting and cloudy days (Figure 4). For example, the averaged predictive errors of CASA
174
model was about -1.12 g C m-2 day-1 at overcastting days, however, the predictive errors was 0.15
175
g C m-2 day-1 at clear days. Previous of studies have found that increased fraction of diffuse
176
radiation at the cloudy days enhanced the plant photosynthesis (Gu et al., 2002, 2003; Urban et al.,
177
2007; Alton et al., 2007). For example, Gu et al (2003) reported increase in diffuse radiation
178
because of volcanic aerosols alone enhanced noontime photosynthesis of a deciduous forest by 23%
179
in 1992 and 8% in 1993. This finding contributed to the temporary increase of terrestrial
180
ecosystems carbon sink after the eruption of Mount Pinatubo (15.1ºN, 121.4ºE) on 15 June 1991
181
(Ciais et al., 1995; Bousquet et al., 2000; Battle et al., 2000). Besides volcanic, cloud reduces the
182
global solar radiation but increases the relative proportion of diffuse radiation at the Earth surface
183
too.
184
<<Figure 4>>
185
It is a case that increased fraction of diffuse radiation can be the cause of changes in
186
many atmospheric factors such as temperature, moisture, and latent heating etc. These factors all
187
have direct or indirect influences on terrestrial ecosystem carbon dynamics (Gu et al., 1999).
188
Therefore some researchers emphasized decreases in the respiration of sunlit leaves due to
189
reduced leaf temperature and vapor pressure deficit (Baldocchi, 1997). However, direct impacts of
190
diffuse radiation had been found: (1) Diffuse radiation results in higher light use efficiencies by
9
191
plant canopies (Gu et al., 2002; Alton et al., 2007). (2) Diffuse radiation penetrates to lower depths
192
of the canopy more efficiently than does the direct radiation (Matsuda et al., 2004). This increased
193
the potential leaf area available for photosynthesis. (3) An increase of blue/red light ratio may lead
194
to higher photosynthesis rates per unit leaf area with increasing faction of diffuse radiation
195
(Matsuda et al., 2004).
196
Among the investigated seven LUE models, CFlux is only one model to integrate the
197
impacts of diffuse radiation on plant photosynthesis (Turner et al., 2006). CFlux model assumed a
198
maximum potential LUE at the overcast condition and minimum LUE value at the clear days, and
199
then a linear decreased trend with cloud cover (Turner et al., 2006). The results showed the simply
200
linear equation successfully reflected the impacts of diffuse radiation. The latest study developed a
201
two-leaf light use efficiency (TL-LUE) model based on the MOD17 algorithm, which separates
202
the canopy into sunlit and shaded leaf groups and calculates GPP separately for them with
203
different maximum light use efficiencies (He et al., 2013). The newly developed TL-LUE model
204
shows lower sensitivity to sky conditions than the MOD17 algorithm.
205
3.2 Comparison of interannual variability
206
The interannual variability of vegetation production exists at different scales from global,
207
regional to plot/stand levels and has been linked with the interannual variability of global carbon
208
balance as well as atmospheric CO2 concentration (Goulden et al., 1996; Bousquet et al., 2000;
209
Zhao et al., 2010). Previous study indicated interannual variability of the global vegetation
210
production is a major driver of the interannual CO2 growth rate (Zhao et al., 2010). Therefore,
211
understanding the cause and degree of interannual variability is important for both ecological
10
212
213
theory and global carbon cycle.
In this study, the ability of simulating interannual variability was investigated at 100
214
sites with more than three-year observations. The results showed the poor ability of seven models
215
to identify the interannual variability. The correlation coefficient (R2) of standard deviation
216
between simulations and observations through all sites ranged from 0.14 to 0.54, and EC-LUE,
217
CFlux and CASA models showed the highest R2 (Figure 5).
218
<<Figure 5>>
219
Our result confirmed a previous study that modeling interannual variation in GPP has
220
proven challenging (Richardson et al., 2007). Keenan et al (2012) assessed the performance of 16
221
terrestrial biosphere models and 3 remote sensing products against long-term measurements of
222
biosphere-atmosphere CO2 exchange made with eddy-covariance flux towers at 11 forested sites
223
in North America. The results showed none of the models consistently reproduce observed
224
interannual variability within measurement uncertainty. Compared with the process-based models,
225
the remote sensing GPP products performed comparably to the average process-based model when
226
assessed against interannual variability (Keenan et al., 2012). Although the response of terrestrial
227
ecosystems to mean climatic drivers is relatively well captured, sensitivity to the impact of
228
variability in climatic drivers may not be, leading to the accumulation of high frequency model
229
error (Dietze et al., 2011) over longer time scales (Schwalm et al., 2010). Although estimates of
230
GPP based on remote sensing have been used to evaluate process-based models, results herein
231
suggest that estimates of interannual variability from both approaches are subject to similar
232
magnitudes of error (Poulter et al., 2011).
233
The possible causes of the errors for modeling interannual variability of GPP could be:
11
234
(1) Incomplete integration of environmental regulations. Most of LUE models only integrate the
235
impacts of temperature, water and radiation. Few LUE models consider the impacts of stand age,
236
phenology and CO2 fertilization on vegetation production, however, previous of studies have
237
showed the significant regulations of those environmental variables to vegetation production
238
(White et al., 1999; DeLucia and Thomas, 2000). (2) Limited understands on the key
239
physiological processes. Braswell et al. (1997) showed that climate induced physiological changes
240
are greater than the direct effect of climatic variability on the carbon cycle. Hui et al. (2003) used
241
a sum-of-squares approach to separate interannual variability in carbon cycle into four different
242
sources: functional change, interannual climatic variability, seasonal climatic variability and
243
random error, and seasonal climatic variability can explain mostly interannual variability in
244
carbon cycles. However, current LUE models do not integrated any responses of physiological
245
processes to environment changes.
246
3.3 Comparison of model structure
247
Pairwise comparison showed higher correlations of PLUE (i.e. PAR×fPAR×LUEmax)
248
among seven models compared with those of GPP simulations (Fig.6). For example, the
249
correlations of CASA model and other six models on PLUE ranged from 0.75 to 0.96, however,
250
correlations of GPP simulations were found from 0.37 to 0.43 (Fig.6). Generally, the pairwise
251
comparison of PLUE and GPP simulations can essentially indicate the contributions of fPAR and
252
environment regulation scalars to the differences of GPP simulations. The result implied larger
253
contribution of environment stress equations on model differences compared with the fraction of
254
PAR absorbed by the vegetation canopy.
12
255
256
<<Figure 6>>
Within LUE models, the fraction of solar radiation intercepted by terrestrial vegetation
257
(fPAR) is calculated from remote sensing data, which is a critical variable for estimating vegetation
258
production. Various methods were developed for calculating fPAR among the LUE models. At the
259
CASA model, fPAR was calculated as a linear function of the simple ratio (SR), which was derived
260
from NDVI (Potter et al., 1993). CFlux and MODIS models directly utilized MODIS-fPAR
261
products. Other four models calculated fPAR based on vegetation index. CFix and EC-LUE
262
models used linear equation with NDVI to estimate fPAR (Veroustraete et al., 2002; Yuan et al.,
263
2007), and VPM and VPRM directly used EVI to indicate fPAR (Xiao, et. al., 2004). CASA, CFix
264
and EC-LUE applied the equation with NDVI to calculate fPAR and showed the high consistence
265
through most sites (Fig.6). Low correlations were observed among MODIS-fPAR, EVI and
266
NDVI-based fPAR (Fig.6)
267
The second pairwise comparison indicated the larger impacts of water stress equation on
268
GPP simulations than that of temperature stress equation (Fig.7). On average, the correlations of
269
RLUEtem between two models were 0.80±0.08, while the mean value of correlations of RLUEwater
270
was 0.65±0.12. In generally, temperature and water response equations were the important two
271
down-regulation factors for LUE models.
272
<<Figure 7>>
273
We compared the consistence of water stress derived from seven models within sites.
274
Water stress scalars of seven models were separated into three levels: drought, normal and wet.
275
The results showed the more than 50% water stress levels were not consistent among models
276
(Table2). The largest inconsistence of water levels was found between MODIS and EC-LUE/CFix
13
277
with 65% differences (Table 2). Moreover, results also showed large friction of inconsistent
278
identifying drought and wet conditions. As shown at Table 2, on average among all models, more
279
than 15% days were wrongly identified between drought and wet days (Table 2). Comparison of
280
water stress equations among LUE models showed the substantial differences through the almost
281
all sites (Fig.8). Moisture availability of CASA showed the largest correlation with that of
282
EC-LUE model through the sites, and the weak relationship with that of MODIS-GPP. Among
283
other four models, moisture availability hardly showed significant relationship. For example,
284
EC-LUE models indicated the site with similar moisture availability, on the contrary, MODIS-GPP
285
product showed the large difference (Fig.8). The moisture response curve of VPM and VPRM is
286
unique and showed the different spatial distribution compare with those of other models.
287
<<Figure 8>>
288
Significant differences of model performance were found among seven LUE models at
289
the three water stress condition (i.e. drought, normal and wet) (Fig.9). Except CFlux, VPRM and
290
VPM, other four models showed low R2 at the drought days. For example, CFix model explained
291
about 50±25% of the variation of GPP estimated at the wet conditions averagely, however, only
292
explained 22±15% variation at the drought days (Fig. 9). Moreover, no consistent predictive errors
293
were found under the different water stresses among the various models. CASA model tended to
294
underestimate GPP at the drought days, on the contrary, CFix model showed obvious
295
overestimation of GPP (Fig. 9).
296
<<Figure 9>>
297
298
Defining a function for quantifying the control of moisture availability on plant
photosynthesis has long been a challenge. The effects of water on plant photosynthesis have been
14
299
estimated as a function of soil moisture, evapotranspiration friction and water vapor pressure
300
deficit (VPD) in a number of LUE models (Field et al., 1995; Prince and Goward, 1995; Running
301
et al., 2000). For instance, in the EC-LUE model, water stress was estimated using the ratio of
302
actual evapotranspiration to net shortwave radiation energy. This ratio was considered to be a very
303
good indicator of soil or vegetation moisture conditions because decreasing amounts of energy
304
partitioned into latent heat flux suggests a stronger moisture limitation (Kurc and Small, 2004;
305
Zhang et al., 2004; Suleiman and Crago, 2004). Other of models, such as VPM and VPRM, used a
306
satellite-derived water index (Land Surface Water Index) to estimate the seasonal dynamics of
307
water stress (Xiao et al., 2004).
308
These variables, which have been used into LUE models, had their weaknesses. For
309
example, it is difficult to characterize soil moisture conditions over large areas from either
310
modeling or remote sensing. This limits the predictive power of any spatial GPP model that relies
311
on soil moisture. VPD is not a good indicator of the spatial heterogeneity of soil moisture
312
conditions across the landscape (e.g., slope versus valley) and it is not likely to be linearly related
313
to soil water availability for which it is often used as a proxy. Moreover, evapotranspiration
314
friction needs an ET model for simulating ecosystem evapotranspiration, and any uncertainties
315
within ET models will reduce the model performance of LUE models.
316
4. Summary
317
We evaluated seven satellite-driven light use efficiency models against 155 eddy
318
covariance sites globally including six major biomes. All seven models showed similar model
319
performance over the vegetation types. The best model performance was observed at deciduous
15
320
broadleaf forests and mixed forests, intermediate at grasslands and evergreen needleleaf forests,
321
lowest at evergreen broadleaf forests. From the spatial respective, CFlux and EC-LUE models
322
showed higher correlations between site-averaged GPP observations and simulations, and were
323
represented with higher performance to simulate interannual variability of GPP. fPAR, temperature
324
curves and water stress equations largely differed among the seven LUE models. Comparably,
325
water stress equations differed largestly which was the major cause for GPP simulations
326
difference.
327
328
329
Acknowledgments
This study was supported by the National Natural Science Foundation of China
330
(41201078), the National High Technology Research and Development Program of China (863
331
Program) (2013AA122003), Program for New Century Excellent Talents in University
332
(NCET-12-0060) and the Fundamental Research Funds for the Central Universities.
333
334
References
335
Alton, P. B., North, P. R. & Los, S. O. (2007). The impact of diffuse sunlight on canopy light-use
336
efficiency, gross photosynthetic product and net ecosystem exchange in three forest biomes.
337
Global Change Biology, 14(3), 776–787.
338
Baldocchi, D. (1997). Measuring and modeling carbon dioxide and water vapor exchange over a
339
temperate broad-leaved forest during the 1995 summer drought. Plant, Cell & Environment.
340
20(9), 1108–1122.
16
341
Battle, M., Bender, M. L., Tans, P. P., White, J. W. C., Ellis, J. T., Conway, T., et al (2000). Global
342
carbon sinks and their variability inferred from atmospheric O2 and δ13C. Science, 287,
343
2467–2470.
344
Baldocchi, D., Falge, E., Gu, L., Olson, R., Hollinger, D., Running, S., et al. (2001). FLUXNET: a
345
new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water
346
vapor and energy flux densities. Bulletin of the American Meteorological Society, 82(11),
347
2415–2435.
348
Bunn, A. G., & Goetz, S. J. (2006). Trends in satellite-observed circumpolar photosynthetic
349
activity from 1982 to 2003: the influence of seasonality, cover type, and vegetation density.
350
Earth Interactions, 10(12), 1–19.
351
Bousquet, P., Peylin, P., Ciais, P., Le Quéré, C., Friedlingstein, P. & Tans, P. P. (2000). Regional
352
changes in carbon dioxide fluxes of land and ocean since 1980. Science, 290, 1342–1345.
353
Beer, C., Reichstein, M., Tomelleri, E., Ciais, P., Jung, M., Carvalhais, N., et al. (2010). Terrestrial
354
gross carbon dioxide uptake: global distribution and covariation with climate. Science, 329,
355
834–838.
356
357
358
Braswell, B. H., Schimel, D. S., Linder, E. & Moore, B. (1997). The response of global terrestrial
ecosystems to interannual temperature variability. Science, 278, 870–872.
Cramer, W., Bondeau, A., Moore, B., Churkina, C., Nemry, B., Ruimy, A., et al. (1999).Comparing
359
global models of terrestrial net primary productivity (NPP): overview and key results. Global
360
Change Biology, 5(S1), 1−15.
361
362
Chen, W., Chen, J. M., Price D. T. & Cihlar J. (2002). Effects of stand age on net primary
productivity of boreal black spruce forests in Ontario, Canada. Canadian Journal for Forest
17
363
364
Research, 32(5), 833–842.
Canadell, J. G., Kirschbaum, M. U. F., Kurz, W. A., Sanz, M. J., Schlamadinger, B., & Yamagata,
365
Y. (2007). Factoring out natural and indirect human effects on terrestrial carbon sources and
366
sinks. Environmental Science & Policy, 10(4), 370–384.
367
368
369
Chapin III, F. S., Matson, P. A., Mooney, H. A. (2012). Principles of terrestrial ecosystem ecology.
Springer, 2nd edition.
Ciais, P., P. P. Tans, M. Trolier, J. W. C. White, & R. J. Francey. (1995). A large northern
370
hemisphere terrestrial CO2 sink indicated by the 13C/12C ratio of atmospheric CO2, Science,
371
269, 1098–1102.
372
Coops, N. C., Waring, R. H., Law, B. E. (2005). Assessing the past and future distribution and
373
productivity of ponderosa pine in the Pacific Northwest using a process model, 3-PG.
374
Ecological Modelling, 183(1), 107–124.
375
376
377
DeLucia, E. & Thomas, R. (2000). Photosynthetic responses to CO2 enrichment of four hardwood
species in a forest understory. Oecologia, 122(1), 11–19.
Dietze, M. C., Vargas, R., Richardson, A. D., Stoy, P. C., Barr, A. G., Anderson, R. S., et al. (2011).
378
Characterizing the performance of ecosystem models across time scales: a spectral analysis of
379
the North American Carbon Program site-level synthesis. Journal of Geophysical Research:
380
Biogeosciences, 116(G4), 2005–2012.
381
Field, C. B., Jackson, R. B., & Mooney, H. A. (1995). Stomatal responses to increased CO2:
382
implications from the plant to the global scale. Plant, Cell & Environment, 18(10), 1214–1225.
383
Friedl, M. A., Sulla-Menashe, D., Tan, B., Schneider, A., Ramankutty, N., Sibley, A., et al. (2010).
384
MODIS Collection 5 global land cover: algorithm refinements and characterization of new
18
385
386
datasets. Remote Sensing of Environment, 114(1), 168–182.
Gu, L., Baldocchi, D., Verma, S. B., Black, T. A., Vesala, T., Falge, E. M., et al. (2002).
387
Advantages of diffuse radiation for terrestrial ecosystem productivity. Journal of geophysical
388
research: Atmospheres, 107(D6), ACL 2-1–ACL 2-23.
389
Gu, L. H., Baldocchi, D., Wofsy, S. C., Munger, J. W., Michalsky, J. J., Urbanski, S. P., et al.
390
(2003). Response of a deciduous forest to the mount Pinatubo eruption: enhanced
391
photosynthesis. Science, 299, 2035–2038.
392
Gu, L. H., Fuentes, J. D., Shugart, H. H., Staebler, R. M., & Black, T. A. (1999). Responses of net
393
ecosystem exchanges of carbon dioxide to changes in cloudiness: results from two North
394
American deciduous forests. Journal of Geophysical Research: Atmospheres, 104(D24),
395
31421–31434.
396
Goulden, M. L., Munger, J. W., Fan, S. M., Daube, B. C., & Wofsy, S.C. (1996). Exchange of
397
carbon dioxide by a deciduous forest: response to interannual climate variability. Science, 271,
398
1576–1578.
399
He MZ, Ju WM, Zhou YL, Chen JM, He HL, Wang SQ, Wang HM, Guan DX, Yan JH, Li YN,
400
Hao YB, Zhao FH. (2013). Development of a two-leaf light use efficiency model for
401
improving the calculation of terrestrial gross primary productivity. Agricultural and Forest
402
Meteorology, 173, 28-39.
403
404
405
406
Hui, D., Luo, Y., & Katul, G. (2003). Partitioning interannual variability in net ecosystem
exchange between climatic variability and functional change. Tree physiology, 23(7), 433–442.
Huntzinger, D. N., Post, W. M., Wei, Y., Michalak, A. M., West, T. O., Jacobson, A. R., et al.
(2012). North American Carbon Program (NACP) regional interim synthesis: terrestrial
19
407
408
biospheric model intercomparison. Ecological Modelling, 232, 144–157.
Heinsch, F. A., Zhao, M., Running, S. W., Kimball, J. S., Nemani, R. R., Davis, et al. (2006).
409
Evaluation of remote sensing based terrestrial productivity from MODIS using regional tower
410
eddy flux network observations. IEEE Transactions on Geoscience and Remote Sensing, 44(7),
411
1908–1925.
412
Keenan, T. F., Baker, I., Barr, A., Ciais, P., Davis, K., Dietze, M., et al. (2012). Terrestrial
413
biosphere model performance for inter-annual variability of land-atmosphere CO2 exchange.
414
Global Change Biology, 18(6), 1971-1987.
415
Knapp, A. K., Fay, P. A., Blair, J. M., Collins, S. L., Smith, M. D., Carlisle, J. D., et al. (2002).
416
Rainfall variability, carbon cycling, and plant species diversity in a mesic grassland. Science,
417
298, 2202–2205.
418
Kositsup, B., Montpied, P., Kasemsap, P., Thaler, P., Améglio, T., & Dreyer, E. (2009).
419
Photosynthetic capacity and temperature responses of photosynthesis of rubber trees (Hevea
420
brasiliensis Müll. Arg.) acclimate to changes in ambient temperatures. Trees, 23(2), 357–365.
421
Kurc, S. A., & Small, E. E. (2004). Dynamics of evapotranspiration in semiarid grassland and
422
shrubland ecosystems during the summer monsoon season, central New Mexico. Water
423
Resources Research, 40(9), 305, doi:10.1029/2004WR003068.
424
425
426
427
428
King, D. A., Turner, D. P., & Ritts, W. D. (2011). Parameterization of a diagnostic carbon cycle
model for continental scale application. Remote Sensing of Environment, 115(7), 1653-1664.
Landsberg, J.J. (1986). Physiological ecology of forest production. Academic Press, London,
165-178.
Landsberg, J. J., & Sands, P. (2010). Physiological ecology of forest production: principles,
20
429
processes and models (Vol. 4). Academic Press, London, 165–178.
430
Landsberg, J. J., & Waring, R. H. (1997). A generalised model of forest productivity using
431
simplified concepts of radiation-use efficiency, carbon balance and partitioning. Forest
432
Ecology and Management, 95(3), 209–228.
433
Melillo, J. M., Kicklighter, D. W., McGuire, A. D., Peterjohn, W. T., & Newkirk, K. M. (1995).
434
Global change and its effects on soil organic carbon stocks. In Dahlem Conference
435
Proceedings. (pp. 175–189). New York: John Wiley and Sons.
436
Matsuda, R., Ohashi-Kaneko, K., Fujiwara, K., Goto, E., & Kurata, K. (2004). Photosynthetic
437
characteristics of rice leaves grown under red light with or without supplemental blue light.
438
Plant & cell physiology, 45(12), 1870-1874.
439
Mahadevan, P., Wofsy, S. C., Matross, D. M., Xiao, X. M., Dunn, A. L., Lin, J. C., et al. (2008). A
440
satellite-based biosphere parameterization for net ecosystem CO2 exchange: Vegetation
441
Photosynthesis and Respiration Model (VPRM). Global Biogeochemical Cycles, 22(2), 1–17.
442
Poulter, B., Ciais, P., Hodson, E., Lischke, H., Maignan, F., Plummer, S., et al. (2011). Plant
443
functional type mapping for earth system models. Geoscientific Model Development, 4(4),
444
993–1010.
445
446
447
Prince, S. D., & Goward, S. N. (1995). Global primary production: a remote sensing approach.
Journal of biogeography, 22(4/5), 815–835.
Potter, C. S., Randerson, J. T., Field, C. B., Matson, P. A., Vitousek, P. M., Mooney, H. A., et al.
448
(1993). Terrestrial ecosystem production: a process model based on global satellite and surface
449
data. Global Biogeochemical Cycles, 7(4), 811–841.
450
Qi, Y., Xu, M., & Wu, J., (2002). Temperature sensitivity of soil respiration and its effects on
21
451
452
ecosystem carbon budget: nonlinearity begets surprise. Ecological Modelling. 153, 131–142.
Raupach, M. R., Canadell, J. G., & Le Quéré, C. (2008). Anthropogenic and biophysical
453
contributions to increasing atmospheric CO2 growth rate and airborne fraction. Biogeosciences
454
Discuss, 5(4), 2867–2896.
455
Richardson, A., Hollinger, D., Aber, J., Ollinger, S., and Braswell, B. (2007). Environmental
456
variation is directly responsible for short but not long term variation in forest atmosphere
457
carbon exchange. Global Change Biol., 13, 788-803.
458
Rossini, M., Cogliati, S., Meroni, M., Migliavacca, M., Galvagno, M., Busetto, L., et al. (2012).
459
Remote sensing-based estimation of gross primary production in a subalpine grassland.
460
Biogeosciences, 9(7), 2565–2584.
461
Ruimy, A., Kergoat, L., & Bondeau, A. (1999). Comparing global models of terrestrial net primary
462
productivity (NPP): analysis of differences in light absorption and light-use efficiency. Global
463
Change Biology, 5(S1), 56–64.
464
Running, S. W., Nemani, R. R., Heinsch, F. A., Zhao, M., Reeves, M., & Hashimoto, H. (2004). A
465
continuous satellite-derived measure of global terrestrial primary production. Bioscience,
466
54(6), 547–560.
467
Ruimy, A., Saugier, B., & Dedieu, G. (1994). Methodology for the estimation of terrestrial net
468
primary production from remotely sensed data. Journal of Geophysical Research:
469
Atmospheres, 99(D3), 5263–5283.
470
Running, S. W., Thornton, P. E., Nemani, R., & Glassy, J. M. (2000). Global terrestrial gross and
471
net primary productivity from the Earth Observing System. Methods in ecosystem science,
472
44–57.
22
473
474
475
476
Schimel, D. (2007). Carbon cycle conundrums. Proceedings of the National Academy of Sciences,
104(47), 18353–18354.
Suleiman, A., Crago, R., (2004). Hourly and daytime evapotranspiration from grassland using
radiometric surface temperatures. Agronomy Journal. 96, 384–390.
477
Schaefer, K., Schwalm, C. R., Williams, C., Arain, M. A., Barr, A., Chen, J. M., et al. (2012). A
478
model-data comparison of gross primary productivity: Results from the North American
479
Carbon Program site synthesis. Journal of Geophysical Research: Biogeosciences, 117(G3),
480
1–15.
481
Schwalm, C. R., Williams, C. A., Schaefer, K., Arneth, A., Bonal, D., Buchmann, N., et al. (2010).
482
Assimilation exceeds respiration sensitivity to drought: A FLUXNET synthesis. Global
483
Change Biology, 16(2), 657–670.
484
Turner, D. P., Ritts, W. D., Styles, J. M., Yang, Z., Cohen, W. B., Law, B. E., et al. (2006). A
485
diagnostic carbon flux model to monitor the effects of disturbance and interannual variation in
486
climate on regional NEP. Tellus B, 58(5), 476−490.
487
Turner, D. P., Urbanski, S., Bremer, D., Wofsy, S. C., Meyers, T., Gower, S. T., et al. (2003). A
488
cross-biome comparison of daily light use efficiency for gross primary production. Global
489
Change Biology, 9(3), 383–395.
490
Urban, O., JANOUŠ, D., Acosta, M., CZERNÝ, R., MarkovA, I., NavrATil, M.,et al. (2007).
491
Ecophysiological controls over the net ecosystem exchange of mountain spruce stand:
492
comparison of the response in direct vs. diffuse solar radiation. Global Change Biology, 13(1),
493
157–168.
494
Veroustraete, F., Sabbe, H., & Eerens, H. (2002). Estimation of carbon mass fluxes over Europe
23
495
496
using the C-Fix model and Euroflux data. Remote Sensing of Environment. 83(3), 376–399.
Viña, A., & Gitelson, A. A. (2005). New developments in the remote estimation of the fraction of
497
absorbed photosynthetically active radiationin crops, Geophysical Research Letters, 32(17),
498
L17403, doi:10.1029/2005GL023647.
499
Walter-Shea, E. A., Privette, J., Cornell, D., Mesarch, M. A., & Hays, C. J. (1997). Relations
500
between directional spectral vegetation indices and leaf area and absorbed radiation in alfalfa.
501
Remote Sensing of Environment, 61(1), 162–177.
502
White, M. A., Running, S. W. & Thornton, P. E. (1999). The impact of growing-season length
503
variability on carbon assimilation and evapotranspiration over 88 years in the eastern US
504
deciduous forest. International Journal of Biometeorology, 42(3), 139–145.
505
Xiao, X. M., Zhang, Q. Y., Braswell, B., Urbanski, S., Boles, S., Wofsy, S., et al. (2004). Modeling
506
gross primary production of temperate deciduous broadleaf forest using satellite images and
507
climate data. Remote Sensing of Environment. 91(2), 256–270.
508
Yuan, W. P., Liu, S. G., Yu, G. R., Bonnefond, J. M., Chen, J. Q., Davis, K., et al. (2010). Global
509
estimates of evapotranspiration and gross primary production based on MODIS and global
510
meteorology data. Remote Sensing of Environment. 114(7), 1416–1431.
511
Yuan, W. P., Liu, S. G., Zhou, G. S., Zhou, G. Y., Tieszen, L. L., Baldocchi, D., et al. (2007).
512
Deriving a light use efficiency model from eddy covariance flux data for predicting daily gross
513
primary production across biomes. Agricultural and Forest Meteorology, 143(3), 189–207.
514
Zhang, Y. Q., Liu, C. M., Yu, Q., Shen, Y. J., Kendy, E., Kondoh, A., et al. (2004). Energy fluxes
515
and the Priestley-Taylor parameter over winter wheat and maize in the North China Plain.
516
Hydrological Processes, 18(12), 2235–2246.
24
517
518
Zhao, M., & Running, S. W. (2010). Drought-induced reduction in global terrestrial net primary
production from 2000 through 2009. Science, 329, 940–943.
25
519
Table 1 Calibrated model parameter values for seven models at the two parameterization
520
schemes.
Vegetation Type
Parameter
CSH
DBF
EBF
ENF
GRA
MIF
CASA
ɛ0
0.62±0.20
1.22±0.43
0.87±0.15
0.85±0.18
0.78±0.17
1.04±0.36
0.55±0.26
1.27±0.07
1.33±0.15
1.09±0.06
1.14±0.09
1.16±0.13
ɛmax
1.12±0.28
3.07±0.25
3.02±0.18
2.29±0.12
2.53±0.15
2.53±0.25
ɛcs
0.66±0.05
1.17±0.07
1.12±0.10
0.95±0.05
1.08±0.08
1.05±0.13
Tmin
-13.76±6.54
-4.35±2.13
-14.46±4.35
-14.20±2.40
-20.12±7.06
-14.26±4.17
Tmax
13.04±8.79
13.08±3.23
20.00±0.00
8.25±1.82
9.55±5.87
13.79±3.74
VPDmin
0.33±0.34
0.11±0.01
0.12±0.07
0.11±0.00
0.12±0.01
0.12±0.02
VPDmax
3.42±0.58
2.99±0.32
2.56±0.32
2.79±0.13
3.23±0.47
2.44±0.32
1.28±0.40
1.71±0.19
1.70±0.11
1.85±0.20
1.59±0.41
1.72±0.31
ɛ0
0.66±0.28
1.77±0.19
1.68±0.10
1.36±0.08
1.52±0.16
1.64±0.22
Tmin
-13.76±6.54
-4.35±2.13
-14.46±4.35
-14.20±2.40
-20.12±7.06
-14.26±4.17
Tmax
13.04±8.79
13.08±3.23
20.00±0.00
8.25±1.82
9.55±5.87
13.79±3.74
VPDmin
0.33±0.34
0.11±0.01
0.12±0.07
0.11±0.02
0.12±0.01
0.12±0.02
VPDmax
3.42±0.58
2.99±0.32
2.56±0.32
2.79±0.13
3.23±0.47
2.44±0.32
1.25±0.43
2.11±0.11
2.17±0.16
2.17±0.10
1.92±0.12
2.03±0.24
ɛ0
4.42±1.88
8.63±1.14
10.88±2.17
14.89±2.10
7.87±1.08
10.16±3.04
PAR0
4.47±1.22
3.15±0.44
2.37±0.75
1.61±0.24
3.07±0.52
2.50±0.81
CFix
ɛ0
CFlux
EC-LUE
ɛ0
MODIS
VPM
ɛ0
VPRM
521
522
523
CSH, DBF, EBF, ENF, GRA, MIF: calibrated parameter values within shrubland, deciduous broadleaf
forest, evergreen broadleaf forest, evergreen needleleaf forest, grassland, mixed forest respectively.
Parameters were introduced at the Supplemental Online Material.
26
Table 2 Comparison of water stress levels derived from seven LUE models
CASA
CFix/EC-LUE
CFlux
MODIS
VPRM/VPM
-
0.49±0.12
0.15±0.28
0.57±0.15
0.56±0.13
CFix/EC-LUE
0.09±0.04
-
0.52±0.12
0.65±0.11
0.58±0.11
CFlux
0.12±0.05
0.06±0.06
-
0.50±0.15
0.54±0.12
MODIS
0.16±0.09
0.19±0.09
0.18±0.09
-
0.61±0.12
VPRM/VPM
0.16±0.10
0.22±0.14
0.13±0.09
0.17±0.10
-
CASA
27
524
Figure caption
525
Figure 1 Model performance of seven Light Use Efficiency Models with calibrated parameters at
526
various vegetation types.
527
Figure 2 Observed vs. the simulated GPP over the 155 EC sites with calibrated parameters. The
528
long dash line is 1:1 line and the solid line is linear regression line.
529
Figure 3 Percentage of eddy covariance sites with higher domination coefficient and lower RMSE
530
for individual model compared the mean values of seven models.
531
Figure 4 The model performance at clear, cloudy and overcast days for seven models.
532
Figure 5 Correlation between standard deviations of simulated interannual variability of GPP.
533
Figure 6 Pairwise comparisons of correlations of GPP simulations and potential light energy use
534
(PLUE) among seven LUE models.
535
Figure 7 Pairwise comparisons of correlations of realized light energy use only considering
536
temperature stress (RLUEtem) and water stress (RLUEwater) among seven LUE modes.
537
Figure 8 Comparison of water stress curves among seven LUE models.
538
Figure 9 The model performance at drought, normal and wet days for seven models.
539
28
540
541
Figure 1 Model performance of seven Light Use Efficiency Models with calibrated parameters at
542
various vegetation types.
29
543
544
Figure 2 Observed vs. the simulated GPP over the 155 EC sites with calibrated parameters. The
545
long dash line is 1:1 line and the solid line is linear regression line.
30
546
547
Figure 3 Percentage of eddy covariance sites with higher domination coefficient and lower RMSE
548
for individual model compared the mean values of seven models.
31
549
550
Figure 4 The model performance at clear, cloudy and overcast days for seven models.
32
551
552
Figure 5 Correlation between standard deviations of simulated interannual variability of GPP.
33
553
554
Figure 6 Pairwise comparisons of correlations of GPP simulations and potential light energy use
555
(PLUE) among seven LUE models.
34
556
557
Figure 7 Pairwise comparisons of correlations of realized light energy use only considering
558
temperature stress (RLUEtem) and water stress (RLUEwater) among seven LUE modes.
35
559
560
Figure 8 Comparison of water stress curves among seven LUE models.
36
561
562
Figure 9 The model performance at drought, normal and wet days for seven models.
37
Download