Next Article in Journal
SAHARA: A Simplified AtmospHeric Correction AlgoRithm for Chinese gAofen Data: 1. Aerosol Algorithm
Previous Article in Journal
Stay-Green and Associated Vegetative Indices to Breed Maize Adapted to Heat and Combined Heat-Drought Stresses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Data-Driven Assessment of Biosphere-Atmosphere Interaction Impact on Seasonal Cycle Patterns of XCO2 Using GOSAT and MODIS Observations

1
Key Laboratory of Digital Earth Science, Institute of Remote Sensing and Digital Earth, Chinese Academy of Sciences, Beijing 100094, China
2
College of Resources and Environment, University of Chinese Academy of Sciences, Beijing 100049, China
3
Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA
*
Authors to whom correspondence should be addressed.
Remote Sens. 2017, 9(3), 251; https://doi.org/10.3390/rs9030251
Submission received: 20 October 2016 / Accepted: 6 March 2017 / Published: 8 March 2017

Abstract

:
Using measurements of the column-averaged CO2 dry air mole fraction (XCO2) from GOSAT and biosphere parameters, including normalized difference vegetation index (NDVI), enhanced vegetation index (EVI), leaf area index (LAI), gross primary production (GPP), and land surface temperature (LST) from MODIS, this study proposes a data-driven approach to assess the impacts of terrestrial biosphere activities on the seasonal cycle pattern of XCO2. A unique global land mapping dataset of XCO2 with a resolution of 1° by 1° in space, and three days in time, from June 2009 to May 2014, which facilitates the assessment at a fine scale, is first produced from GOSAT XCO2 retrievals. We then conduct a statistical fitting method to obtain the global map of seasonal cycle amplitudes (SCA) of XCO2 and NDVI, and implement correlation analyses of seasonal variation between XCO2 and the vegetation parameters. As a result, the spatial distribution of XCO2 SCA decreases globally with latitude from north to south, which is in good agreement with that of simulated XCO2 from CarbonTracker. The spatial pattern of XCO2 SCA corresponds well to the vegetation seasonal activity revealed by NDVI, with a strong correlation coefficient of 0.74 in the northern hemisphere (NH). Some hotspots in the subtropical areas, including Northern India (with SCA of 8.68 ± 0.49 ppm on average) and Central Africa (with SCA of 8.33 ± 0.25 ppm on average), shown by satellite measurements, but missed by model simulations, demonstrate the advantage of satellites in observing the biosphere–atmosphere interactions at local scales. Results from correlation analyses between XCO2 and NDVI, EVI, LAI, or GPP show a consistent spatial distribution, and NDVI and EVI have stronger negative correlations over all latitudes. This may suggest that NDVI and EVI can be better vegetation parameters in characterizing the seasonal variations of XCO2 and its driving terrestrial biosphere activities. We, furthermore, present the global distribution of phase lags of XCO2 compared to NDVI in seasonal variation, which, to our knowledge, is the first such map derived from a completely data-driven approach using satellite observations. The impact of retrieval error of GOSAT data on the mapping data, especially over high-latitude areas, is further discussed. Results from this study provide reference for better understanding the distribution of the strength of carbon sink by terrestrial ecosystems and utilizing remote sensing data in assessing the impact of biosphere–atmosphere interactions on the seasonal cycle pattern of atmospheric CO2 columns.

Graphical Abstract

1. Introduction

Understanding the interactions between terrestrial ecosystem and atmospheric CO2 column is significant to understand the variations of greenhouse gases in our atmosphere. The terrestrial biosphere is a large and persistent carbon sink [1]. Every year about 1/3 of the anthropogenic CO2 emission from fossil fuels combustion, cement production, and land-use change is absorbed by the terrestrial ecosystems [2]. The CO2 uptake and release induced by terrestrial biosphere activities essentially contribute to the seasonal cycle of atmospheric CO2 column [3], which is, however, weakly related to fossil emission [2] and ocean fluxes [4]. The CO2 seasonal cycle amplitude (SCA) is mainly determined by the difference between the dominant CO2 uptake during the growing seasons and CO2 release during the autumn and winter seasons [5,6]. To understand the underlying mechanisms, surface measurements and satellite observations have been widely used in analyzing the seasonality of atmospheric CO2 [7,8,9,10,11,12,13].
CO2 data from current surface monitoring networks have been crucial in gaining important insights of the mechanisms of CO2 seasonal variations [7,8,9,10]. Seasonal cycle of CO2 at high northern hemisphere (NH) latitudes has been found to be dominantly induced by terrestrial activities, especially within tundra and boreal forest biomes using data from long-term surface measurements and atmosphere transport models [10]. Variations in the strength of carbon sinks over North America can be well monitored by measurements from ground sites [7]. An enhanced strength of the seasonal exchange of CO2, which is substantially larger than current model simulations in northern extratropical land ecosystems, especially in boreal forest, can be found from ground-based and aircraft-based observations [9]. Climate–vegetation–carbon cycle feedback at high latitudes and the latitudinal gradient of the increasing CO2 SCA can be precisely observed by a total carbon column-observing network (TCCON) [8]. However, in global scale, it is still difficult to capture the full picture with detailed variations of biosphere–atmospheric CO2 interaction by current ground measurement networks. With the advantages of global coverage and high-density measurements, satellite observations have the potential to improve our understandings in both global and local scales.
Results from validation with TCCON measurements show that XCO2 retrievals from the Greenhouse gases Observing SATellite (GOSAT) well capture the XCO2 seasonal cycle [14]. Spatial distributions of SCA, derived from gridded data over 10° by 10° in space, are consistent among satellite observations from GOSAT and the SCanning Imaging Absorption spectroMeter for Atmospheric CHartographY (SCIAMACHY) and model simulations from Monitoring Atmospheric Composition and Climate (MACC 13.1) and CarbonTracker (CT2013b) at the global scale [15]. In addition, the latitudinal gradient of biosphere–atmosphere CO2 interactions has been revealed with satellite observations [11,13,16]. However, geographic distribution of the magnitude of biosphere–atmospheric CO2 exchange, which drives the atmospheric CO2 seasonal cycle, and its underlying mechanisms are still highly uncertain [17,18]. Analysis at finer scale of the interaction using biophysical parameters and XCO2 retrievals from satellite observations is still lacking. One of the challenges is the irregular distribution and gaps of the XCO2 retrievals in space and time, which result from certain constraints, such as cloud coverage and satellite observation modes [7]. As a solution, a global mapping XCO2 dataset derived from GOSAT XCO2 retrievals, which has been demonstrated to be accurate in characterizing the spatio-temporal variations of XCO2 at the global scale [19,20], can be made to investigate the detailed seasonal cycle pattern of XCO2.
On the other hand, satellite observations of terrestrial biosphere have been widely used in studying the terrestrial vegetation activities resulting in terrestrial carbon uptake and release. Vegetation indices and parameters, such as normalized difference vegetation index (NDVI), enhanced vegetation index (EVI), leaf area index (LAI), and gross primary production (GPP), indicate the strength of photosynthesis activity related to terrestrial CO2 uptake [21,22,23], while land surface temperature (LST) has an impact on the respiration rate and evaporation of terrestrial biosphere [24]. For a better understanding of where and how strong the biosphere’s seasonal activities influence the seasonal cycle pattern of XCO2, this study proposes a data-driven approach to assess the impacts from terrestrial biosphere–atmosphere interactions on the seasonal cycle pattern of XCO2 using satellite retrievals of XCO2 observed by GOSAT and parameters related to terrestrial biosphere activities observed by the Moderate Resolution Imaging Spectroradiometer (MODIS) from June 2009 to May 2014. To use the satellite XCO2 retrievals, which are irregularly distributed in space and time and have many gaps, a data-driven mapping approach based on spatio-temporal geostatistics is applied to fill the gaps and generate a gridded, mapping XCO2 dataset [19,20] with a resolution of 1° by 1° in space, and three days in time (hereafter referred to as GM-XCO2). The satellite observations of terrestrial biosphere, including NDVI, EVI, LAI, GPP, and LST from MODIS, are collected and re-gridded into the same spatial resolution with the mapping XCO2 dataset.
In Section 2, we describe all the used data and methods. Results of XCO2 SCA from GOSAT, correlation analysis of seasonal cycle between XCO2 and terrestrial biosphere parameters, and seasonal cycle phase lags between XCO2 and biosphere activities indicated by NDVI are presented in Section 3. Discussions on the uncertainty of GM-XCO2, and possible impact of LST on seasonal variation of XCO2 on tropical land are presented in Section 4, followed by conclusions in Section 5.

2. Materials and Methods

2.1. Datasets Used in the Study

Satellite observations from GOSAT and MODIS, model simulations from CarbonTracker, and ground-based measurements for validation from TCCON for 5 years from 1 June 2009 to 31 May 2014 were collected. The specifications of the used datasets are shown in Table 1. To highlight the differences of seasonal cycle over various land cover types, we selected nine different regions of interest where there are distinctive correlations of seasonal cycle between XCO2 and terrestrial biosphere parameters. Figure 1 shows the locations of these regions overlapped with land cover map. Table 2 gives detailed specifications of these regions. In addition, XCO2 data from 12 TCCON sites (shown in Figure 1) and model simulations are obtained to investigate the uncertainty of GM-XCO2 datasets used in this study.

2.1.1. Global Land Mapping XCO2 Data Derived from GOSAT Retrievals

Global land mapping dataset of XCO2, spanning from 1 June 2009 to 31 May 2014, is generated by filling gaps of ACOS-GOSAT Level 2 XCO2 retrievals (v3.5), downloaded from the NASA data portal [32], using the mapping approach developed by Zeng et al. [19,20] based on spatio-temporal geostatistics. A detail description and accuracy assessment of the global mapping method can be found in Zeng et al. [20]. Unlike Zeng et al. [20], in this study, when modeling the spatio-temporal trend during the mapping process, we use the original ACOS-GOSAT XCO2 data instead of model simulations. In this way, the method makes full use of information inside the data and is therefore a completely data-driven mapping approach. The resulted mapping XCO2 data, GM-XCO2, are gridded dataset and regular in space and time with spatial resolution of 1° by 1° and temporal resolution of 3 days. To screen out data with high uncertainty, mapping data with prediction standard deviation [19] larger than 2.0 ppm are not used in this study. Accuracy assessment of this mapping data is presented in Section 4.1. With high resolution in space and time, this unique global land mapping dataset of XCO2 make it possible for both spatial and temporal comparison analysis at a fine scale with other satellite-observed parameters. As described in Zeng et al. [19], the mapping result and its data uncertainty from this data-driven gap-filling method rely heavily on the original XCO2 retrievals. Therefore, any possible retrieval artifacts due to retrieval errors will be propagated into the mapping data. More description on the mapping data uncertainty will be discussed in Section 4.

2.1.2. Terrestrial Biosphere Parameters from MODIS

The satellite observations related to terrestrial biosphere activities and land surface temperature used are obtained from MODIS onboard the Terra or Aqua. These data include MYD13C2-NDVI, MYD13C2-EVI, MCD15A2-LAI, MOD17A2-GPP and MOD11C3-LST. NDVI and EVI are the vegetation indices positively related to the amount of the biomass and strength of vegetation photosynthesis [28], LAI indicates the state of plant growth [30], GPP represents the ability of plants to capture carbon [29], and LST affects the soil and plant respiration rates and ecosystem photosynthesis [33]. It has been shown that, in the tropical areas, the net effect of increasing LST is increasing atmospheric CO2 [34]. A detailed description of all the used datasets is shown in Table 1.

2.1.3. TCCON

TCCON, a global network of ground-based Fourier transform spectrometers established for measuring atmospheric XCO2 and other trace gases [35]. It has a high accuracy with approximately 0.25% error in the retrieved XCO2, which has been extensively used for validation and calibration of satellite observations, including ACOS-GOSAT retrievals [36]. In this study, the 2014 release version of the TCCON data (GGG2014) for 12 sites (with at least 2 years of coincidental measurements and within the mapping area; as shown in Figure 1) is used. Measurements within the period from June 2009 to May 2014 were selected for comparison. The 12 selected sites include Park Falls [37,38], Orleans [39], Garmisch [40], Karlsruhe [41], Bremen [42], Bialystok [43,44], Tsukuba [45,46], Lamont [47], JPL [48], Saga [49], Darwin [50], and Wollongong [51].

2.1.4. XCO2 Simulations from Carbon-Tracker

Model simulations of CO2 used for comparison are from CarbonTracker CT2015 [27], which has been adopted for comparison in several previous studies [11,13,20]. In this study, the XCO2 value is calculated from the CO2 profile data within the local time around 13:00, using a pressure-averaged method as described in Conner et al. [52]. The XCO2 data from CT2015 is hereafter referred to as CT-XCO2. When comparing GM-XCO2 and CT-XCO2, the difference due to averaging the kernel effect [49] is not considered here, since the difference of XCO2 between applying and not applying the averaging kernel smoothing on model simulations is less than 0.1% [53].

2.2. Calculation of Seasonal Cycle Amplitude of XCO2

The SCA of XCO2 is determined by CO2 uptake and release by terrestrial biosphere. The temporal variation of XCO2 can be divided into two components, long-term yearly increases mainly caused by anthropogenic emissions [3] and inherent seasonal cycle mainly resulted from photosynthesis and respiration by the terrestrial biosphere [23]. Fitting formulas for the XCO2 seasonal cycle have been proposed in many different forms [13,14,20,54,55]. In this study, we use the typical one, adopted by Keeling et al. [54] and Thoning et al. [55], to statistically fit the seasonal cycle and yearly increases of XCO2, which is given by the Equation (1) below,
x ( t ) = k 0 + k 1 × t + i = 1 2 ( a i cos ( 2 π i t ) + b i sin ( 2 π i t ) )
where k 0 is the XCO2 at the start time, k 1 is the temporal growth rate of XCO2, and the annual harmonic functions, i = 1 2 ( a i cos ( 2 π i t ) + b i sin ( 2 π i t ) ) , describe the seasonal variation of XCO2. The harmonic functions include an annual seasonal cycle and semi-annual seasonal cycle [56]. We fit Equation (1) to the five-year time series of GM-XCO2. However, the inter-annual variations in the seasonal cycle shape is not included in the fitted curves. In this study, the yearly increase of XCO2 in each grid is removed to obtain the seasonal GM-XCO2 data (hereafter referred to as dXCO2), which are used in analyzing the seasonality in the following sections. The SCA is calculated as the difference between maximum and minimum value of dXCO2 from five years GM-XCO2 data from Jun 2009 to May 2014. An example of fitting GM-XCO2 using Equation (1) in a selected grid (46.5°N, 111.5°E) is shown in Figure 2.

2.3. Correlating Analysis of Seasonal Variation between XCO2 and Biosphere Parameters

The seasonal variation of XCO2 is driven by the seasonal imbalance between CO2 uptake and release related to ecosystem photosynthesis and respiration [13], which are the outcomes of the underlying biosphere–atmospheric interactions. However, it is still not clear how much of the seasonal variation of XCO2 can be explained by surface biosphere parameters at global scale with fine resolutions, especially from a data-driven perspective. We use the GM-XCO2 dataset and terrestrial biosphere parameters, including NDVI, EVI, LAI, GPP, and LST data, and implement correlation analyses on their seasonal cycles. Pearson correlation coefficients (PCCs) in seasonal variation between XCO2 and each of the parameters (NDVI, EVI, LAI, GPP, and LST) for the terrestrial biosphere are calculated for each grid using monthly mean values, respectively. XCO2 data less than 350 ppm are excluded; NDVI and EVI larger than 0 and valid LST, GPP, and LAI were used. The spatial distribution of PCC between the monthly mean of dXCO2 and the biosphere parameters was used to quantify the impact of biosphere activities on the seasonal cycle pattern of XCO2.

3. Results

3.1. Spatial Pattern of Seasonal Cycle Amplitude of XCO2 and NDVI

XCO2 SCA is an effective indicator of the strength of seasonal variation of XCO2, which is related to the increase and drawdown of CO2 in the atmosphere [13]. The spatial variation of SCA has been shown by Lindqvist et al. [14] using ground measurements from sparse TCCON stations. Further work by Kulawik et al. [15] using satellite retrievals has shown the global variation of SCA in 10° by 10° grids based on a traditional binning method, which, however, does not consider the detailed variation information within each grid. Based on our global land mapping data, Figure 3 shows the global land distribution of XCO2 SCA derived from seasonal cycle of GM-XCO2 from June 2009 to May 2014 using Equation (1). This is, to our knowledge, the first reported global map of CO2 SCA at fine scale derived from real measurements. The goodness-of-fit evaluated by coefficient of determination (R2) between mapping and fitting XCO2, as described in Section 2.2, is shown in Figure S1. We can see that R2 for all fittings is larger than 0.9, indicating that the fittings well capture the seasonal cycle of GM-XCO2. This global map of XCO2 SCA clearly shows the latitude gradient decrease from north to south that is consistent with previous studies using ground measurements and binned satellite observations [15,35,57]. In the southern hemisphere (SH), the XCO2 SCA are mostly lower than 4.0 ppm except some areas in Southern Africa, where the amplitude is above 4.0 ppm. It can be seen from this SCA map, moreover, that the range border of greater than 8 ppm extended over the high latitudes mostly with boreal forest in the NH. Moreover, the tropical and subtropical areas in India and Central Africa show abnormal SCA values larger than 8 ppm, which have not been seen previously and can be detected from our mapping dataset with fine spatial details.
To investigate how biosphere activities drive the spatial pattern of XCO2, we correlated the XCO2 SCA, as shown in Figure 3, with NDVI SCA, an indicator of the strength of vegetation photosynthesis and its resulting carbon uptake. Corresponding to Figure 3, Figure 4 presents the NDVI SCA overlapped with amplitude contour lines of GM-XCO2 shown in Figure 3, which is derived in the same way as XCO2 SCA for the same period. Comparing Figure 3 with Figure 4, it can be seen that the overall spatial distribution of XCO2 SCA shows a very similar pattern as that of NDVI, with a high correlation coefficient of 0.69 for global land and 0.74 for the NH land. Both the SCAs of XCO2 and NDVI generally decrease with latitude from north to south. Especially, the amplitude of XCO2 has the largest value ranging from 8.0 ppm up to 12 ppm over high latitude in the NH, corresponding well to the highest amplitude of NDVI in these areas, which is likely caused by the strong CO2 uptake during the growing season by boreal forest (Figure 1) [13,58]. The SCA of XCO2 changes mostly from 4.0 ppm to 8.0 ppm over the mid-latitude area in the NH. It is remarkable to find that large amplitudes up to 8 ppm are presented over the Northern Indian area (8.68 ± 0.49 ppm) and some hotspots in Central Africa (8.33 ± 0.25 ppm), whereas the NDVI SCAs in these areas are relatively low. These areas, partly shown as CnI and GcA in Figure 1, are probably due to the CO2 release from soil respiration of terrestrial biosphere in Northern India region or the large wild fires in the central Africa area. In addition, the temporal variation of anthropogenic emissions on the XCO2 SCA in Northern India may partly contribute to these SCA anomalies. Further discussion is presented in Section 4.3.

3.2. Correlation between XCO2 and Biosphere Parameters

In this section, correlation analyses are conducted, respectively, between seasonal variations of GM-XCO2 data and several biophysical parameters associated with biosphere activities, including NDVI, EVI, LAI and GPP, to explore the underlying biosphere–atmospheric coupling process that drives the XCO2 seasonality. Figure 5 shows the global distribution of the Pearson Correlation Coefficients (PCCs) with a significant level (p-value < 0.01) between monthly-averaged dXCO2 and NDVI, EVI, LAI, and GPP, respectively, from June 2009 to May 2014 in each grid. Due to the lack of available data, the tropical forests in Amazon and Southeast Asia and the Tibetan Plateau are not considered here.
All of the PCCs between seasonal cycles of XCO2 and seasonal variations of NDVI, EVI, LAI, and GPP, shown in Figure 5, present similar spatial patterns and are mostly negatively correlated, except for the bare land areas (shown in Figure 1). Especially, the seasonal cycle of XCO2 show stronger and significant negative correlations with the biosphere parameters over Northeast Asia around 45°N than those over the other areas. Among the indices of NDVI, EVI, LAI, and GPP, XCO2 shows stronger negative correlation with NDVI and EVI than those with LAI and GPP globally. In the NH high-latitude region, the correlation coefficients are −0.69, −0.68, −0.66, and −0.59 for NDVI, EVI, LAI, and GPP, respectively. They are −0.74, −0.69, −0.64, and −0.67 in the NH low-latitude region. It is known that LAI and GPP are the vegetation parameters derived from semi-empirical biophysical models, which likely induce more uncertainty than NDVI and EVI that are directly retrieved from MODIS-observed radiance. Moreover, it can be seen that the PCCs between XCO2 and LAI are slightly weaker than those from other indices in the low latitudes, which is likely because LAI is an indicator of vegetation biophysical structure rather than biochemical activities [59,60]. The PCCs from EVI, moreover, are slightly higher than that from NDVI over Northeast China. This difference may be related to the saturation effect of NDVI in high biomass regions, while EVI data does not become saturated as easily as NDVI by using the non-saturated NIR band for high biomass regions [61].
We selected six regions of interest with a PCC higher than 0.5 distributed across multiple latitudes as shown in Figure 6 to investigate the correlation of seasonal variations of XCO2 and NDVI from June 2009 to May 2014. These six regions of interest are shown in Figure 1, including FnC: forest in Northern Canada covered by 63% of forest, FeR: forest in Eastern Russia covered by 86% of forest, CwR: crop in Western Russia covered by 70% of cropland, CnC: crop in Northern China covered by 58% of cropland with mixed 18% grassland and shrub-land, GcA: grass in Central Africa covered by 36% of grassland mixed with 30% of forest, and CnI: crop in Northern India covered by 89% of cropland.
It can be seen from Figure 6 that XCO2 decreases sharply when NDVI increases quickly during the growing season in summer. A stronger negative correlation in the forest region (−0.68 in FeR) compared to the cropland region (−0.50 in CwR) at the same latitude zone can be found. PCCs between seasonal variations of XCO2 and NDVI is up to 0.8 over GcA, CnI, and CnC. The rainy season over GcA boosts the grass growth and promote the CO2 uptake. Strong correlation presented in CnI and CnC, imply that the seasonal crop farming from planting, growing to harvesting may have a large impact on the seasonal variation of CO2 uptake and release. Moreover, from Figure 6, we can see that there is an offset between the time of XCO2 troughs and NDVI peaks. The time delay, as lags of month, over the six regions of interest are also indicated in Figure 6. Comparing with phase of NDVI, the phase of XCO2 is generally delayed by a lag of about 0–2 months, which generally agrees with results from Table 3 in [62] by Olsen et al., which shows the delayed time between seasonal amplitudes of XCO2 and boundary layer CO2 to be within 2 months [62], with generally a greater difference in high latitude regions compared to low ones in NH. This offset is probably due to the time delay in XCO2 response to terrestrial biosphere activities associated with the mass of CO2 in the column and vertical mixing and convection of CO2 from the surface to the column profile [58,62,63]. In this study, we define the time delay to be the lags of months between XCO2 and NDVI that make their negative correlation the strongest. We can see the delayed time over the three regions (FnC, FeR, and CwR) in high latitude regions are over one month, and larger than the other three regions at low latitudes (CnC, GcA, and CnI), probably indicating a relatively slow vertical mixing due to very large changes of the CO2 mass at high latitudes with strong biosphere absorption. The global pattern of this phase delay effect is elaborated in Section 3.3.

3.3. Global Pattern of Phase Delay Effect of XCO2 to NDVI

While almost impossible for surface measurements, the GM-XCO2 and NDVI data in this study allow us to investigate the global distribution of phase delay effect in terms of lags of months. The results are shown in Figure 7a, which is the averaged phase lag for five years of GM-XCO2 and NDVI data. This map of phase delay effect, to our knowledge, is the first such map derived from a completely data-driven approach using satellite observations.
Generally, the phase delay in southern hemisphere is either zero or one, probably due to the much smaller change of CO2 mass in the column that needs time to be well mixed in column. In the NH mid- and high-latitude regions, interestingly, the pattern of the phase delay is similar to the seasonal cycle pattern derived from GM-XCO2 shown in Figure 3 and from NDVI shown in Figure 4. The larger amplitudes, such as North America and North Eurasia boreal forest regions, correspond to a smaller time delay, suggesting that the strength of vertical mixing is much stronger in these areas, even though the biosphere–atmosphere exchange in these areas is also large. As shown in [64] by Huang et al., the strength of vertical mixing in boreal and arctic regions is not as strong as that in the tropics, which may lead to stronger phase delay effect in high latitude regions. Moreover, XCO2 has a large footprint, which is supposed to have contributions from different oceanic and land regions. However, satellite observations of XCO2, such as from ACOS-GOSAT, are more sensitive to CO2 variations within the boundary layer [65], which allows us to explore the biosphere–atmosphere interactions by correlating seasonal cycles of satellite-observed XCO2 and surface biosphere parameters.
Figure 7b shows the PCC between seasonal variation of XCO2 and NDVI if the phase delay effect is excluded. We can see that, over most of the regions, the PCC is stronger than −0.8, particularly for most areas in North America where PCCs are around −0.6 before excluding the effect. In addition, the negative correlation over some area of boreal forest, most areas of grassland and forest in Central Africa have been increased to stronger than −0.9. These results show that the correlation between seasonal cycles of XCO2 and NDVI may be underestimated if this phase lag effect is not removed. Furthermore, it suggests that this phase lag effect associated with atmospheric transport should be considered when assessing the impact of biosphere–atmosphere interactions on seasonal cycle patterns of XCO2.

4. Discussion

4.1. Uncertainty of XCO2 Global Land Mapping Dataset

4.1.1. Evaluation Using Cross-Validation

Cross-validation is widely used in accuracy assessments for statistical models [20,66]. In this study, we used cross-validation based on the Monte Carlo sampling technique adopted by Zeng et al. [20]. The Monte Carlo method is carried out by randomly sampling 5% of the data that will be left out in cross-validation, and repeating this process 100 times. Please refer to Zeng et al. [20] for technical details. From the outputs of cross-validation, two datasets, the predicted dataset and the original observation dataset, are obtained. The distribution of observed values and predicted values of XCO2 is shown in Figure 8. We can see the predicted XCO2 well agrees with observed XCO2 with high R2 of 0.91, indicating a very strong correlation between the original ACOS-GOSAT data and the predictions. The mean absolute prediction error (MAPE), which is the averaged prediction error, is 0.85 ppm, indicating that the averaged absolute error for each prediction is less than 1 ppm. Specifically, 70% of the prediction errors are less than 1 ppm. These statistics from cross-validation study suggest that the used mapping method based on spatio-temporal geostatistics is effective and precise in generating global land mapping of XCO2 from original retrievals. The uncertainty in the mapping XCO2 data comes from a combination of the uncertainty from the mapping method and the uncertainty propagated from satellite retrieval error. This latter source of data uncertainty is discussed in the following section by comparing with TCCON.

4.1.2. Verification with TCCON

TCCON has been widely applied in verification of XCO2 from satellite observations [14,20,36]. In this study, we perform a comparison between TCCON XCO2 data for 12 sites and the XCO2 time series reconstructed from the global land mapping XCO2 data. For each TCCON site, we use all available data within the period from June 2009 to May 2014. The averaging kernel, as described in [67], for comparing ACOS-GOSAT and TCCON data is used for this comparison. Time series of TCCON XCO2, GM-XCO2, and ACOS-GOSAT XCO2 within 500 km of every TCCON site are shown in Figure S2. Table 3 shows the statistical results in terms of the number of coincident data pairs, bias, and SCA difference between the mapping dataset and TCCON. The SCA of TCCON XCO2 shown in Table 3 is derived in the same way as SCA of GM-XCO2, as shown in Figure S2 which depicts the seasonal variations of XCO2 from ACOS-GOSAT retrievals, mapping data, and TCCON measurements.
Figure S2 compares the seasonal variation of observed and mapping XCO2 and TCCON measurements. We can see that the mapping XCO2 data go through the original satellite observations and the mapping data well capture the seasonal variations of the observations. From Table 3, we can see that the overall absolute difference of SCA between GM-XCO2 and TCCON is 0.4 ppm. The SCAs from GM-XCO2 are systematically shallower than that from TCCON for all sites in continental Europe, which is consistent with conclusions from previous studies [14,15] of XCO2 SCA using original GOSAT-ACOS retrievals, due to retrieval biases that were evaluated in previous studies [14,15]. The SCAs of GM-XCO2 are well consistent to that of TCCON (SCA difference < 0.7 ppm) at Park Falls, Lamont and Wollongong with long-term measurement that covers all the five years as shown in Figure S1. The largest absolute different SCA are 1.15 ppm in Darwin, 1.75 ppm at Orleans, and 1.38 ppm at Karlsruhe. It should be noted that Darwin is a coastal site of SH, whose SCA is small and has a combined contribution from local, oceanic, and adjacent land regions. From the comparison results, we can see that the mapping dataset can generally capture the seasonal characteristics of XCO2. However, the mapping result and its data uncertainty from this data-driven gap-filling method rely heavily on the original XCO2 retrievals. As a result, the retrieval artifacts on original XCO2 retrievals are propagated into the mapping dataset [19], primarily leading to the differences between the seasonal cycles of GM-XCO2 and TCCON. As shown in Chevallier et al. [68], evidence of systematic retrieval errors can be found over dark surfaces of the high-latitude lands, particularly in the boreal region. These retrieval errors, which propagate into the mapping data, may play a role in the difference when compared to TCCON as shown in Table 3 and may lead to underestimation of XCO2 SCA in the boreal region as shown in Figure 3. However, these possible errors (about 1 ppm for Continental Europe) are not likely to have a substantial impact on the SCA patterns, as we can see in Figure 3. The retrieval errors on XCO2 retrievals may have a smaller impact on the correlation analysis between seasonal cycles between GM-XCO2 and biosphere parameters, as shown throughout this paper, since it is mostly the phase of seasonal cycles that determine the PCC value instead of absolute values of XCO2.

4.1.3. Comparison of Seasonal Cycle between GM-XCO2 and Model Simulations

As suggested by previous studies [11,27,53,58], model simulations can well reproduce the large-scale features of atmospheric CO2. On the other hand, satellite observations make direct measurements of XCO2 and, therefore, be able to capture the local variability and a more detailed characteristic of XCO2 variation compared to model simulations. Therefore, we compare the SCA of GM-XCO2 to that of simulated XCO2 from CarbonTracker in order to evaluate the satellite measurements and the mapping dataset. Figure 9 shows the global distribution of XCO2 SCA derived from the CarbonTracker (CT2015) data, overlapped with the contour lines of GM-XCO2 amplitudes shown in Figure 3. Comparing Figure 3 with Figure 9, it can be clearly seen that the spatial pattern and latitudinal gradient of GM-XCO2 SCA is in good agreement with CT-XCO2, especially in the high-latitude NH areas where the contour lines from GM-XCO2 amplitude match well with the changes of SCA from CT-XCO2. Moreover, the phenomenon of higher SCA over India and Central Africa than surrounding areas can be seen in SCA maps from both GM-XCO2 and CT-XCO2. The SCA difference map of GM-XCO2 and CT-XCO2 is shown in Figure S3. Over the NH high-latitude area, the GM-XCO2 SCA is generally lower than that of CT-XCO2, which agrees with the results, as shown in Table 3, that GM-XCO2 SCA is generally lower than TCCON sites at the same area (Bremen, Karlsruhe, Orleans, Garmisch, and Bialystok). Over the low latitude area, GM-XCO2 SCA is generally higher than that of CT-XCO2, especially over the areas within the overlapped GM-XCO2 contour lines of 8 ppm and 4 ppm (Such as India and Southern Africa). There are two possibilities for these local inconsistencies: (1) the retrieve error of satellite observations may be large over these areas because of the impacts of thin clouds and disturbance of wildfire [57,69]; (2) the uncertainty of simulations by current models [8,70] may be large over these areas because of the sparseness of the ground-based measures and airborne observations used for assimilations to constrain the local CO2 fluxes [20,66]. This comparison shows that satellite observations of XCO2 with appropriate analysis can be used to assess the earth system modeling coupled with carbon flux estimates, as also suggested by Hammerling et al. [71]. We can conclude that the GM-XCO2 SCA is generally consistent with that of CT-XCO2 on a large scale, even though the discrepancies over some local areas need further investigation.

4.2. The Possible Impact of Retrieval Density in High Latitude Regions

As shown in Figure 10, there are large gaps of satellite XCO2 retrievals in high latitude regions, especially during the winter season due to snow cover in these regions and very small signal noise ratio in observation. For analyzing the XCO2 SCA, the availability of retrievals around the peak and trough of the cycle is important. Fortunately, the ACOS-GOSAT XCO2 retrievals cover the peak and trough of the XCO2 cycle, as indicated by CT2015 simulations, suggesting that our data-driven mapping method can generally reproduce the SCAs in these regions from the satellite retrievals. However, at high latitude boreal regions, the ACOS-GOSAT retrievals may also be affected by systematic retrieval errors [14,68], so the SCA values at these high latitude regions, as shown in Figure 3, should be interpreted with caution. Further validations should be carried out with more TCCON data available in the future.

4.3. Possible Impact of LST on Seasonal Variation of XCO2 on Tropical Land

Abnormal results are shown over some regions (e.g., CnI, GcA, FsA, and CeB) in the tropical area (30°N–30°S), where large XCO2 SCA can be seen from both GM-XCO2 and CT-XCO2, even the former is slightly larger than the latter, and high PCCs exist for NDVI and EVI, as shown in Figure 5. These regions are covered by vegetation, such as grass, forest, and crop mixed with bare land, where the interaction between CO2 uptake and release from vegetation and soil respiration likely has impact on the seasonal cycle of CO2. As indicated from previous studies on CO2 emission from soils, the soil respiration in the tropic and subtropical areas is much larger than high latitudes [72,73] and, therefore, likely has non-negligible impact on the XCO2 seasonal pattern. LST, a remotely sensed terrestrial parameter, is effective in depicting the land surface temperature distribution at the global scale with long time-series. LST has been used as one of the critical parameters in estimating seasonal soil respiration over forests in the Midwest USA [74]. Higher surface temperature is associated with stronger soil and plant respiration rates, but also possibly enhanced photosynthesis [33]. However, the net effect of increasing LST in the tropical areas has been shown to result in increasing atmospheric CO2 [34]. In tropical areas, soil and plant respiration contribute a very important part to the seasonal variation of atmospheric CO2. The tropical carbon storage is very vulnerable to temperature change [34]. Therefore, an investigation of the correlation between XCO2 from satellite observation and LST can indicate the possible impact of LST on the seasonal variation of XCO2 in the tropical areas, where the carbon storage has been shown to be sensitive to future temperature variabilities [34].
Figure 11 shows the correlation between seasonal XCO2 and LST variation over the tropical areas. The results interestingly show the seasonal cycle of XCO2 is highly positive correlated with seasonal variation of LST, which is opposite to the negative correlation with NDVI, as shown in Figure 5a. Especially, over Central Africa and Southern Asia, the PCC is generally larger than 0.70, indicating the possible impact of LST on the seasonal cycle of XCO2. Additionally, Figure 12 shows the detailed time series of the selected five regions of interest shown in Figure 1. The seasonal variation of XCO2 is generally consistent with the variation of LST; especially, the variation around peak XCO2 matches well with the variation around peak LST. Coincident with the maximum LST, terrestrial respiration and weakened photosynthesis could contribute to the maximum XCO2. As is shown in Sreenivas et al. [75], XCO2 showed a positive correlation with temperature, and soil respiration and biomass burning may be an important terrestrial carbon source contributing to the CO2 concentration in India. To further investigate the possible influence from anthropogenic emissions on the SCA anomaly, we found that the fossil fuel emission from bottom-up estimation in CnI [76,77] (shown in Figure S4) does show a seasonal cycle, which, however, has much less consistency than LST with XCO2 around the peaks of the XCO2 cycle. This may suggest that the SCA anomaly observed in CnI is partly contributed from fossil fuel emissions but dominantly contributed from LST variations around the spring period. The effect of terrestrial respiration in these regions need to be further investigated and directly observed LST could be an alternative parameter to its further investigation for lack of surface measurements.

5. Conclusions

The seasonal cycle of atmospheric CO2 is mostly caused by CO2 uptake and release from biosphere activity, which is profoundly connected to biosphere fluxes that determine the global net CO2 sink. Satellite observations of XCO2 by GOSAT, correlated with biosphere parameters from MODIS, not only expand the current in situ measurements tremendously, but also provide us with multi-parameter data of biosphere–atmosphere interaction to reveal the impacts of biosphere activity on the seasonal cycle pattern of XCO2. The data-driven analysis of atmospheric CO2 coupled with biosphere parameters has the significant potential to improve our understanding of the global net CO2 sink with massive increases of satellite data. In this study, we introduced a data-driven approach to assess biosphere–atmosphere interaction on the seasonal change of XCO2 through collecting the global data, including XCO2 from GOSAT, vegetation parameters of NDVI, EVI, LAI, and GPP, and land surface temperature from MODIS from the years 2009–2014.
The global land mapping dataset of XCO2 (GM-XCO2), derived from GOSAT XCO2 retrievals, can reveal the spatial distribution of seasonal cycle characteristics of XCO2 directly and at a global scale that were not involved before. Furthermore, GM-XCO2 makes it possible to implement the correlation analyses, at fine scale, of seasonal variation between satellite observed XCO2 and terrestrial biosphere parameters. We found that the seasonal cycle pattern of GM-XCO2 is well corresponding to that from NDVI with high correlation coefficient of 0.69 for global land and 0.74 for the NH land. That indicate the significant influence from vegetation activities on the seasonal change of XCO2. Through the correlation analyses between GM-XCO2 and NDVI, EVI, LAI, and GPP, it is found that NDVI and EVI have better correlations and spatial pattern with the seasonal variation of XCO2 than LAI and GPP for stronger coefficients over all latitude zones. That is consistent with previous studies [78,79] that NDVI and EVI, directly retrieved from MODIS observed radiance, may be better parameters representing biosphere activities for constraining of the terrestrial CO2 flux than LAI and GPP, which are obtained from semi-empirical biophysical models. Using XCO2 and NDVI, moreover, we show the global distribution of time delay effect in terms of lags of months, which, to our knowledge, is the first such map derived from the completely data-driven approach using satellite observations. After removing the phase lags, PCC enhanced nearly −0.2 for most areas, especially for North America, where those were improved from weaker than −0.6 to stronger than −0.8. As a result, the phase lags between seasonal change of XCO2 and the vegetation indices should be considered when assessing the impact of biosphere–atmosphere interaction on the seasonal cycle pattern of XCO2.
Considering the uncertainty of the GM-XCO2, evaluation with cross-validation, comparisons using TCCON measurements and CT-XCO2 was conducted. The method presents a significant high R2 of 0.91 between the predictions and observations in cross-validation. Results of evaluation with TCCON data demonstrate the sufficient accuracy of mapping XCO2 with averaged absolute bias within 2 ppm over all selected 12 TCCON sites, and the similar SCA distribution between GM-XCO2 and CT-XCO2 indicating the availability of GM-XCO2 in capturing the seasonal change of atmospheric CO2 concentration, even though some discrepancy exists in the SCA estimation between GM-XCO2 and CT-XCO2 over low-latitude areas. Considering the strong terrestrial respiration, and its relationship with temperature over low-latitude zones, LST, an important remote sensing parameter, was used for correlation with the seasonal change of XCO2. Fortunately, high positive coefficients are shown in low-latitude area for LST and seasonal change of XCO2, which provide us a potential method of utilizing remote sensing parameters related to the photosynthesis and terrestrial respiration for better understanding the biosphere–atmosphere interaction for lack of surface measurements in the low-latitude areas.
Future studies, including improving the mapping XCO2 data over the tropical areas, comparing the mapping XCO2 with results from other models, and a detailed investigation of the discrepancies between observations and modeling, should be further explored. It is expected that the data-driven assessment developed in this study will be improved as more CO2 observations are becoming available from newly-launched satellite, such as OCO-2, and upcoming satellites, such a as TanSat [80], as well as more remote sensing data on vegetation activities.

Supplementary Materials

The following are available online at www.mdpi.com/2072-4292/9/3/251/s1. Figure S1. The global map of the coefficient of determination (R2) between fitting-XCO2 and GM-XCO2 using Equation (1). R2 in each grid represents the goodness-of-fit of the grid, such as one fitting shown in Figure 2. Figure S2: Temporal variation comparison for the 12 TCCON sites. As shown in these panels, the original ACOS-GOSAT XCO2 retrievals within 500 km of the TCCON site are in gray dots. The TCCON data, smoothed by applying the ACOS-GOSAT averaging kernel, are indicated by blue dots. The data are chosen using coincidence criteria of within ±1 h of the GOSAT overpass time, and a three-day (one time-unit) mean is calculated for the comparison. The predicted TCCON site XCO2 time series using the mapping approach are indicated by the red dots. Figure S3: SCA difference between averaged GM-XCO2 and CT-XCO2 in the mapping area from June 2009 to May 2014, which is calculated using GM-XCO2 shown in Figure 3 minus corresponding values of CT-XCO2 shown in Figure 9. Negative values indicate GM-XCO2 SCA is lower than that of CT-XCO2, and positive values indicate that GM-XCO2 is higher. Figure S4: Monthly mean of detrended XCO2 (dXCO2) and bottom-up estimation of fossil fuel emissions (FFCO2) from ODIAC (Open-source Data Inventory for Anthropogenic CO2) and CDIAC (Carbon Dioxide information Analysis Center) over (a) cropland in Northern India (CnI) and (b) cropland in Northern China (CnC), two regions of interest as shown in Figure 1, from June 2009 to May 2014..

Acknowledgments

This research was supported by the National Research Program on Global Changes and Adaptation: “Big data on global changes: data sharing platform and recognition” (Grant No. 2016YFA0600303, Grant No. 2016YFA0600304). We acknowledge The ACOS-GOSAT v3.5 data were produced by the ACOS/OCO-2 project at the Jet Propulsion Laboratory, California Institute of Technology, and obtained from the ACOS/OCO-2 data archive maintained at the NASA Goddard Earth Science Data and Information Services Center. We also acknowledge the GOSAT Project for acquiring the spectra. CarbonTracker CT2015 results are provided by NOAA ESRL, Boulder, Colorado, USA from the website at http://carbontracker.noaa.gov. Terra MODIS GPP/NPP Product MOD17A2 was downloaded from the University of Montana and Aqua or Terra MODIS NDVI/EVI, LAI products (MYD13C2, MOD11C3) were download from USGS websites (http://e4ftl01.cr.usgs.gov). Land cover products were downloaded from ESA (http://www.esa-ghg-cci.org/). And TCCON data were obtained from the TCCON Data Archive website at http://tccon.ornl.gov/. Bottom-up estimation of fossil fuel emissions (FFCO2) is obtained from ODIAC (Open-source Data Inventory for Anthropogenic CO2: http://www.odiac.org/index.html) and CDIAC (Carbon Dioxide information Analysis Center: http://cdiac.ornl.gov/).

Author Contributions

L.L., Z.H., and Z.-C.Z. conceived and designed the experiments; Z.H. performed the experiments; Z.-C.Z. analyzed the data; N.B., S.Y., and Z.-C.Z. contributed analysis tools; Z.H., L.L., and Z.-C.Z. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, Y.; Birdsey, R.A.; Fang, J.; Houghton, R.; Kauppi, P.E.; Kurz, W.A.; Phillips, O.L.; Shvidenko, A.; Lewis, S.L.; Canadell, J.G.; et al. A large and persistent carbon sink in the world’s forests. Science 2011, 333, 988–993. [Google Scholar] [CrossRef] [PubMed]
  2. Quéré, C.L.; Andres, R.J.; Boden, T.; Conway, T.; Houghton, R.; House, J.I.; Marland, G.; Peters, G.P.; Van der Werf, G.; Ahlström, A. The global carbon budget 1959–2011. Earth Syst. Sci. Data 2013, 5, 165–185. [Google Scholar] [CrossRef] [Green Version]
  3. Stocker, T.; Qin, D.; Plattner, G.; Tignor, M.; Allen, S.; Boschung, J.; Nauels, A.; Xia, Y.; Bex, B.; Midgley, B. Contribution of working group I to the fifth assessment report of the intergovernmental panel on climate change. In Climate Change 2013: The Physical Science Basis; Cambridge University Press: Cambridge, UK, 2013. [Google Scholar]
  4. Doney, S.C.; Lima, I.; Feely, R.A.; Glover, D.M.; Lindsay, K.; Mahowald, N.; Moore, J.K.; Wanninkhof, R. Mechanisms governing interannual variability in upper-ocean inorganic carbon system and air–sea CO2 fluxes: Physical climate and atmospheric dust. Deep Sea Res. Part II Top. Stud. Oceanogr. 2009, 56, 640–655. [Google Scholar] [CrossRef]
  5. Zeng, N.; Zhao, F.; Collatz, G.J.; Kalnay, E.; Salawitch, R.J.; West, T.O.; Guanter, L. Agricultural green revolution as a driver of increasing atmospheric CO2 seasonal amplitude. Nature 2014, 515, 394–397. [Google Scholar] [CrossRef] [PubMed]
  6. Keeling, R.F.; Piper, S.C.; Heimann, M. Global and hemispheric CO2 sinks deduced from changes in atmospheric concentration. Nature 1996, 381, 218–221. [Google Scholar] [CrossRef]
  7. Buermann, W.; Lintner, B.R.; Koven, C.D.; Angert, A.; Pinzon, J.E.; Tucker, C.J.; Fung, I.Y. The changing carbon cycle at Mauna Loa Observatory. Proc. Natl. Acad. Sci. USA 2007, 104, 4249–4254. [Google Scholar] [CrossRef] [PubMed]
  8. Forkel, M.; Carvalhais, N.; Rodenbeck, C.; Keeling, R.; Heimann, M.; Thonicke, K.; Zaehle, S.; Reichstein, M. Enhanced seasonal CO2 exchange caused by amplified plant productivity in northern ecosystems. Science 2016, 351, 696–699. [Google Scholar] [CrossRef] [PubMed]
  9. Graven, H.D.; Keeling, R.F.; Piper, S.C.; Patra, P.K.; Stephens, B.B.; Wofsy, S.C.; Welp, L.R.; Sweeney, C.; Tans, P.P.; Kelley, J.J. Enhanced seasonal exchange of CO2 by northern ecosystems since 1960. Science 2013, 341, 1085–1089. [Google Scholar] [CrossRef] [PubMed]
  10. Randerson, J.T.; Thompson, M.V.; Conway, T.J.; Fung, I.Y.; Field, C.B. The contribution of terrestrial sources and sinks to trends in the seasonal cycle of atmospheric carbon dioxide. Glob. Biogeochem. Cycles 1997, 11, 535–560. [Google Scholar] [CrossRef]
  11. Schneising, O.; Reuter, M.; Buchwitz, M.; Heymann, J.; Bovensmann, H.; Burrows, J.P. Terrestrial carbon sink observed from space: Variation of growth rates and seasonal cycle amplitudes in response to interannual surface temperature variability. Atmos. Chem. Phys. 2014, 14, 133–141. [Google Scholar] [CrossRef]
  12. Welp, L.R.; Patra, P.K.; Rödenbeck, C.; Nemani, R.; Bi, J.; Piper, S.C.; Keeling, R.F. Increasing summer net CO2 uptake in high northern ecosystems inferred from atmospheric inversions and comparisons to remote-sensing NDVI. Atmos. Chem. Phys. 2016, 16, 9047–9066. [Google Scholar] [CrossRef] [Green Version]
  13. Wunch, D.; Wennberg, P.O.; Messerschmidt, J.; Parazoo, N.C.; Toon, G.C.; Deutscher, N.M.; Aleks, G.K.; Roehl, C.M.; Randerson, J.T.; Warneke, T.; et al. The covariation of northern hemisphere summertime CO2 with surface temperature in boreal regions. Atmos. Chem. Phys. 2013, 13, 9447–9459. [Google Scholar] [CrossRef]
  14. Lindqvist, H.; O’Dell, C.W.; Basu, S.; Boesch, H.; Chevallier, F.; Deutscher, N.; Feng, L.; Fisher, B.; Hase, F.; Inoue, M.; et al. Does gosat capture the true seasonal cycle of carbon dioxide? Atmos. Chem. Phys. 2015, 15, 13023–13040. [Google Scholar] [CrossRef]
  15. Kulawik, S.; Wunch, D.; O'Dell, C.; Frankenberg, C.; Reuter, M.; Oda, T.; Chevallier, F.; Sherlock, V.; Buchwitz, M.; Osterman, G.; et al. Consistent evaluation of ACOS-GOSAT, BESD-SCIAMACHY, CarbonTracker, and MACC through comparisons to TCCON. Atmos. Meas. Tech. 2016, 9, 683–709. [Google Scholar] [CrossRef]
  16. Aleks, G.K.; Wennberg, P.O.; Washenfelder, R.A.; Wunch, D.; Schneider, T.; Toon, G.C.; Andres, R.J.; Blavier, J.F.; Connor, B.; Davis, K.J.; et al. The imprint of surface fluxes and transport on variations in total column carbon dioxide. Biogeosciences 2012, 9, 875–891. [Google Scholar] [CrossRef] [Green Version]
  17. Ciais, P.; Dolman, A.J.; Bombelli, A.; Duren, R.; Peregon, A.; Rayner, P.J.; Miller, C.; Gobron, N.; Kinderman, G.; Marland, G.; et al. Current systematic carbon-cycle observations and the need for implementing a policy-relevant carbon observing system. Biogeosciences 2014, 11, 3547–3602. [Google Scholar] [CrossRef] [Green Version]
  18. Schimel, D.; Pavlick, R.; Fisher, J.B.; Asner, G.P.; Saatchi, S.; Townsend, P.; Miller, C.; Frankenberg, C.; Hibbard, K.; Cox, P. Observing terrestrial ecosystems and the carbon cycle from space. Glob. Chang. Biol. 2015, 21, 1762–1776. [Google Scholar] [CrossRef] [PubMed]
  19. Zeng, Z.; Lei, L.; Hou, S.; Ru, F.; Guan, X.; Zhang, B. A regional gap-filling method based on spatiotemporal variogram model of columns. IEEE Trans. Geosci. Remote Sens. 2014, 52, 3594–3603. [Google Scholar] [CrossRef]
  20. Zeng, Z.C.; Lei, L.; Strong, K.; Jones, D.B.A.; Guo, L.; Liu, M.; Deng, F.; Deutscher, N.M.; Dubey, M.K.; Griffith, D.W.T.; et al. Global land mapping of satellite-observed CO2 total columns using spatio-temporal geostatistics. Int. J. Digit. Earth 2017, 10. [Google Scholar] [CrossRef]
  21. Chen, B.; Xu, G.; Coops, N.C.; Ciais, P.; Innes, J.L.; Wang, G.; Myneni, R.B.; Wang, T.; Krzyzanowski, J.; Li, Q.; et al. Changes in vegetation photosynthetic activity trends across the Asia-Pacific region over the last three decades. Remote Sens. Environ. 2014, 144, 28–41. [Google Scholar] [CrossRef]
  22. Inoue, Y.; Peñuelas, J.; Miyata, A.; Mano, M. Normalized difference spectral indices for estimating photosynthetic efficiency and capacity at a canopy scale derived from hyperspectral and CO2 flux measurements in rice. Remote Sens. Environ. 2008, 112, 156–172. [Google Scholar] [CrossRef]
  23. Vetter, P.; Schmid, W.; Schwarze, R. Spatio-temporal statistical analysis of the carbon budget of the terrestrial ecosystem. Stat. Methods Appl. 2016, 25, 1–19. [Google Scholar] [CrossRef]
  24. Huang, N.; Gu, L.; Black, T.A.; Wang, L.; Niu, Z. Remote sensing-based estimation of annual soil respiration at two contrasting forest sites. J. Geophys. Res. Biogeosci. 2015, 120, 2306–2325. [Google Scholar] [CrossRef]
  25. ESA. Available online: http://www.esa-landcover-cci.org/ (accessed on 15 October 2016).
  26. O’Dell, C.W.; Connor, B.; Bösch, H.; O’Brien, D.; Frankenberg, C.; Castano, R.; Christi, M.; Eldering, D.; Fisher, B.; Gunson, M.; et al. The ACOS CO2 retrieval algorithm—Part 1: Description and validation against synthetic observations. Atmos. Meas. Tech. 2012, 5, 99–121. [Google Scholar] [CrossRef]
  27. Peters, W.; Jacobson, A.R.; Sweeney, C.; Andrews, A.E.; Conway, T.J.; Masarie, K.; Miller, J.B.; Bruhwiler, L.M.; Petron, G.; Hirsch, A.I.; et al. An atmospheric perspective on North American carbon dioxide exchange: CarbonTracker. Proc. Natl. Acad. Sci. USA 2007, 104, 18925–18930. [Google Scholar] [CrossRef] [PubMed]
  28. Huete, A.; Didan, K.; van Leeuwen, W.; Miura, T.; Glenn, E. MODIS vegetation indices. In Land Remote Sensing and Global Environmental Change; Springer: Berlin/Heidelberg, Germany, 2011; pp. 579–602. [Google Scholar]
  29. Heinsch, F.A.; Reeves, M.; Votava, P.; Kang, S.; Milesi, C.; Zhao, M.; Glassy, J.; Jolly, W.M.; Loehman, R.; Bowker, C.F. GPP and NPP (MOD17A2/A3) products NASA MODIS land algorithm. In MOD17 User’s Guide; MODIS Land Team: Washington, DC, USA, 2003; pp. 1–57. [Google Scholar]
  30. Yang, W.; Tan, B.; Huang, D.; Rautiainen, M.; Shabanov, N.V.; Wang, Y.; Privette, J.L.; Huemmrich, K.F.; Fensholt, R.; Sandholt, I. MODIS leaf area index products: From validation to algorithm improvement. IEEE Trans. Geosci. Remote Sens. 2006, 44, 1885–1898. [Google Scholar] [CrossRef]
  31. Wan, Z. Collection-5 MODIS Land Surface Temperature Products Users’ Guide; ICESS, University of California: Santa Barbara, CA, USA, 2007. [Google Scholar]
  32. National Aeronautics and Space Administration (NASA). Available online: http://CO2web.jpl.nasa.gov (accessed on 15 October 2016).
  33. Doughty, C.E.; Goulden, M.L. Are tropical forests near a high temperature threshold? J. Geophys. Res. Biogeosci. 2008, 113, 1–12. [Google Scholar] [CrossRef]
  34. Cox, P.M.; Pearson, D.; Booth, B.B.; Friedlingstein, P.; Huntingford, C.; Jones, C.D.; Luke, C.M. Sensitivity of tropical carbon to climate change constrained by carbon dioxide variability. Nature 2013, 494, 341–344. [Google Scholar] [CrossRef] [PubMed]
  35. Wunch, D.; Wennberg, P.O.; Toon, G.C.; Connor, B.J.; Fisher, B.; Osterman, G.B.; Frankenberg, C.; Mandrake, L.; O’Dell, C.; Ahonen, P.; et al. A method for evaluating bias in global measurements of CO2 total columns from space. Atmos. Chem. Phys. 2011, 11, 12317–12337. [Google Scholar] [CrossRef]
  36. Wunch, D.; Toon, G.C.; Blavier, J.F.; Washenfelder, R.A.; Notholt, J.; Connor, B.J.; Griffith, D.W.; Sherlock, V.; Wennberg, P.O. The total carbon column observing network. Philos. Trans. Ser. A Math. Phys. Eng. Sci. 2011, 369, 2087–2112. [Google Scholar] [CrossRef] [PubMed]
  37. Washenfelder, R.A.; Toon, G.C.; Blavier, J.F.; Yang, Z.; Allen, N.T.; Wennberg, P.O.; Vay, S.A.; Matross, D.M.; Daube, B.C. Carbon dioxide column abundances at the Wisconsin tall tower site. J. Geophys. Res. 2006, 111. [Google Scholar] [CrossRef]
  38. Wennberg, P.O.; Roehl, C.; Wunch, D.; Toon, G.C.; Blavier, J.F.; Washenfelder, R.; Aleks, G.K.; Allen, N.; Ayers, J. TCCON Data from Park Falls, Wisconsin, USA, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  39. Warneke, T.; Messerschmidt, J.; Notholt, J.; Weinzierl, C.; Deutscher, N.; Petri, C.; Grupe, P.; Vuillemin, C.; Truong, F.; Schmidt, M.; et al. TCCON Data from Orleans, France, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  40. Sussmann, R.; Rettinger, M. TCCON Data from Garmisch, Germany, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  41. Hase, F.; Blumenstock, T.; Dohe, S.; Gross, J.; Kiel, M. TCCON Data from Karlsruhe, Germany, Release GGG2014R1; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  42. Notholt, J.; Petri, C.; Warneke, T.; Deutscher, N.; Buschmann, M.; Weinzierl, C.; Macatangay, R.; Grupe, P. TCCON Data from Bremen, Germany, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  43. Messerschmidt, J.; Chen, H.; Deutscher, N.M.; Gerbig, C.; Grupe, P.; Katrynski, K.; Koch, F.T.; Lavrič, J.V.; Notholt, J.; Rödenbeck, C.; et al. Automated ground-based remote sensing measurements of greenhouse gases at the Białystok site in comparison with collocated in situ measurements and model data. Atmos. Chem. Phys. 2012, 12, 6741–6755. [Google Scholar] [CrossRef]
  44. Deutscher, N.; Notholt, J.; Messerschmidt, J.; Weinzierl, C.; Warneke, T.; Petri, C.; Grupe, P.; Katrynski, K. TCCON Data from Bialystok, Poland, Release GGG2014R1; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  45. Ohyama, H.; Morino, I.; Nagahama, T.; Machida, T.; Suto, H.; Oguma, H.; Sawa, Y.; Matsueda, H.; Sugimoto, N.; Nakane, H.; et al. Column-averaged volume mixing ratio of CO2 measured with ground-based fourier transform spectrometer at Tsukuba. J. Geophys. Res. 2009, 114. [Google Scholar] [CrossRef]
  46. Morino, I.; Matsuzaki, T.; Shishime, A. TCCON Data from Tsukuba, Ibaraki, Japan, 125hr, Release GGG2014R1; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  47. Wennberg, P.O.; Wunch, D.; Roehl, C.; Blavier, J.F.; Toon, G.C.; Allen, N.; Dowell, P.; Teske, K.; Martin, C.; Martin, J. TCCON Data from Lamont, Oklahoma, USA, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  48. Wennberg, P.O.; Roehl, C.; Blavier, J.F.; Wunch, D.; Landeros, J.; Allen, N. TCCON Data from Jet Propulsion Laboratory, Pasadena, California, USA, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  49. Kawakami, S.; Ohyama, H.; Arai, K.; Okumura, H.; Taura, C.; Fukamachi, T.; Sakashita, M. TCCON Data from Saga, Japan, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  50. Griffith, D.W.T.; Deutscher, N.; Velazco, V.A.; Wennberg, P.O.; Yavin, Y.; Aleks, G.K.; Washenfelder, R.; Toon, G.C.; Blavier, J.F.; Murphy, C.; et al. TCCON Data from Darwin, Australia, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  51. Griffith, D.W.T.; Velazco, V.A.; Deutscher, N.; Murphy, C.; Jones, N.; Wilson, S.; Macatangay, R.; Kettlewell, G.; Buchholz, R.R.; Riggenbach, M. TCCON Data from Wollongong, Australia, Release GGG2014R0; Carbon Dioxide Information Analysis Center: Oak Ridge, TN, USA, 2014.
  52. Connor, B.J.; Boesch, H.; Toon, G.; Sen, B.; Miller, C.; Crisp, D. Orbiting carbon observatory: Inverse method and prospective error analysis. J. Geophys. Res. Atmos. 2008, 113, 1–14. [Google Scholar] [CrossRef]
  53. Cogan, A.J.; Boesch, H.; Parker, R.J.; Feng, L.; Palmer, P.I.; Blavier, J.F.L.; Deutscher, N.M.; Macatangay, R.; Notholt, J.; Roehl, C.; et al. Atmospheric carbon dioxide retrieved from the greenhouse gases observing satellite (GOSAT): Comparison with ground-based TCCON observations and GEOS-Chem model calculations. J. Geophys. Res. Atmos. 2012, 117, 1–17. [Google Scholar] [CrossRef]
  54. Keeling, C.D.; Bacastow, R.B.; Bainbridge, A.E.; Ekdahl, C.A.; Guenther, P.R.; Waterman, L.S.; Chin, J.F.S. Atmospheric carbon dioxide variations at Mauna Loa Observatory, Hawaii. Tellus 1976, 28, 538–551. [Google Scholar]
  55. Thoning, K.W.; Tans, P.P.; Komhyr, W.D. Atmospheric carbon dioxide at Mauna Loa Observatory: 2. Analysis of the NOAA GMCC data, 1974–1985. J. Geophys. Res. Atmos. 1989, 94, 8549–8565. [Google Scholar] [CrossRef]
  56. Jiang, X.; Chahine, M.T.; Li, Q.; Liang, M.; Olsen, E.T.; Chen, L.L.; Wang, J.; Yung, Y.L. CO2 semiannual oscillation in the middle troposphere and at the surface. Glob. Biogeochem. Cycles 2012, 26. [Google Scholar] [CrossRef]
  57. Schneising, O.; Buchwitz, M.; Reuter, M.; Heymann, J.; Bovensmann, H.; Burrows, J.P. Long-term analysis of carbon dioxide and methane column-averaged mole fractions retrieved from SCIAMACHY. Atmos. Chem. Phys. 2011, 11, 2863–2880. [Google Scholar] [CrossRef] [Green Version]
  58. Aleks, G.K.; Wennberg, P.O.; Schneider, T. Sources of variations in total column carbon dioxide. Atmos. Chem. Phys. 2011, 11, 3581–3593. [Google Scholar] [CrossRef]
  59. Gitelson, A.A.; Peng, Y.; Arkebauer, T.J.; Schepers, J. Relationships between gross primary production, green LAI, and canopy chlorophyll content in maize: Implications for remote sensing of primary production. Remote Sens. Environ. 2014, 144, 65–72. [Google Scholar] [CrossRef]
  60. Lausch, A.; Pause, M.; Schmidt, A.; Salbach, C.; Gwillymmargianto, S.; Merbach, I. Temporal hyperspectral monitoring of chlorophyll, LAI, and water content of barley during a growing season. Can. J. Remote Sens. 2013, 39, 191–207. [Google Scholar] [CrossRef]
  61. Huete, A.; Didan, K.; Miura, T.; Rodriguez, E.P.; Gao, X.; Ferreira, L.G. Overview of the radiometric and biophysical performance of the MODIS vegetation indices. Remote Sens. Environ. 2002, 83, 195–213. [Google Scholar] [CrossRef]
  62. Olsen, S.C.; Randerson, J.T. Differences between surface and column atmospheric CO2 and implications for carbon cycle research. J. Geophys. Res. 2004, 109. [Google Scholar] [CrossRef]
  63. Rayner, P.J.; Law, R.M. A Comparison of Modelled Responses to Prescribed CO2 Sources. In CSIRO Division of Atmospheric Research Technical Paper; CSIRO: Canberra, Australia, 1995. [Google Scholar]
  64. Huang, J.; McElroy, M.B. Contributions of the hadley and ferrel circulations to the energetics of the atmosphere over the past 32 years. J. Clim. 2014, 27, 2656–2666. [Google Scholar] [CrossRef]
  65. Aleks, G.K.; Wennberg, P.O.; O’Dell, C.W.; Wunch, D. Towards constraints on fossil fuel emissions from total column carbon dioxide. Atmos. Chem. Phys. 2013, 13, 4349–4357. [Google Scholar] [CrossRef]
  66. Liu, M.; Lei, L.; Liu, D.; Zeng, Z.C. Geostatistical analysis of CH4 columns over monsoon Asia using five years of GOSAT observations. Remote Sens. 2016, 8. [Google Scholar] [CrossRef]
  67. Nguyen, H.; Osterman, G.; Wunch, D.; O’Dell, C.; Mandrake, L.; Wennberg, P.; Fisher, B.; Castano, R. A method for colocating satellite XCO2 data to ground-based data and its application to ACOS-GOSAT and TCCON. Atmos. Meas. Tech. 2014, 7, 2631–2644. [Google Scholar] [CrossRef]
  68. Chevallier, F. On the statistical optimality of CO2 atmospheric inversions assimilating CO2 column retrievals. Atmos. Chem. Phys. 2015, 15, 11133–11145. [Google Scholar] [CrossRef]
  69. Heymann, J.; Schneising, O.; Reuter, M.; Buchwitz, M.; Rozanov, V.V.; Velazco, V.A.; Bovensmann, H.; Burrows, J.P. SCIAMACHY WFM-DOAS XCO2: Comparison with CarbonTracker XCO2 focusing on aerosols and thin clouds. Atmos. Meas. Tech. 2012, 5, 2887–2931. [Google Scholar] [CrossRef]
  70. Chandra, N.; Lal, S.; Venkataramani, S.; Patra, P.K.; Sheel, V. Temporal variations of atmospheric CO2 and CO at Ahmedabad in western India. Atmos. Chem. Phys. 2016, 16, 6153–6173. [Google Scholar] [CrossRef]
  71. Hammerling, D.M.; Michalak, A.M.; O’Dell, C.; Kawa, S.R. Global CO2 distributions over land from the greenhouse gases observing satellite (GOSAT). Geophys. Res. Lett. 2012, 39, 1–6. [Google Scholar] [CrossRef]
  72. Raich, J.W.; Potter, C. Global patterns of carbon dioxide emission from soils. Glob. Biogeochem. Cycles 1995, 9, 23–36. [Google Scholar] [CrossRef]
  73. Raich, J.W.; Potter, C.; Bhagawati, D. Interannual variability in global soil respiration, 1980–94. Glob. Chang. Biol. 2002, 8, 800–812. [Google Scholar] [CrossRef]
  74. Huang, N.; Gu, L.; Niu, Z. Estimating soil respiration using spatial data products: A case study in a deciduous broadleaf forest in the Midwest USA. J. Geophys. Res. Atmos. 2014, 119, 6393–6408. [Google Scholar] [CrossRef]
  75. Sreenivas, G.; Mahesh, P.; Subin, J.; Kanchana, A.L.; Rao, P.V.N.; Dadhwal, V.K. Influence of meteorology and interrelationship with greenhouse gases (CO2 and CH4) at a suburban site of India. Atmos. Chem. Phys. 2016, 16, 3953–3967. [Google Scholar] [CrossRef]
  76. Oda, T.; Maksyutov, S. A very high-resolution (1 km×1 km) global fossil fuel CO2 emission inventory derived using a point source database and satellite observations of nighttime lights. Atmos. Chem. Phys. 2011, 11, 543–556. [Google Scholar] [CrossRef]
  77. Andres, R.J.; Gregg, J.S.; Losey, L.; Marland, G.; Boden, T.A. Monthly, global emissions of carbon dioxide from fossil fuel consumption. Tellus B 2011, 63, 309–327. [Google Scholar] [CrossRef]
  78. Frank, A.B.; Karn, J.F. Vegetation indices, CO2 flux, and biomass for northern plains grasslands. J. Range Manag. 2003, 56, 382–387. [Google Scholar] [CrossRef]
  79. Aalto, T.; Ciais, P.; Chevillard, A.; Moulin, C. Optimal determination of the parameters controlling biospheric CO2 fluxes over Europe using eddy covariance fluxes and satellite NDVI measurements. Tellus B 2004, 56, 93–104. [Google Scholar] [CrossRef]
  80. Liu, Y.; Yang, D.; Cai, Z. A retrieval algorithm for Tansat XCO2 observation: Retrieval experiments using GOSAT data. Chin. Sci. Bull. 2013, 58, 1520–1523. [Google Scholar] [CrossRef]
Figure 1. Regions of interest selected from global land region from 45°S to 60°N overlapped with land-cover data in 2009 obtained from Europe Space Agency (ESA, [25]). Nine different regions of interest are outlined as red rectangles, including Forest in Northern Canada (FnC), Forest in Eastern Russia (FeR), Cropland in Western Russia (CwR), Cropland in Northern China (CnC), Cropland in Northern India (CnI), Cropland in Eastern Brazil (CeB), Grassland in Central Africa (GcA), Mixed forest in Southern Africa (FsA) and Shrub-land in Northern Australia (SnA). Six regions in the Northern Hemisphere, including FnC, CwR, FeR, CnC, CnI, and GcA, are used for correlation analysis of seasonal cycle between NDVI and XCO2 in Section 3.2, while five regions, including CnI, GcA, CeB, FsA, and SnA, are used for correlation analysis of seasonal variations between LST and XCO2 in Section 4.2. Twelve selected sites of TCCON are marked with black dots and used for evaluating mapping data in Section 4.1.2.
Figure 1. Regions of interest selected from global land region from 45°S to 60°N overlapped with land-cover data in 2009 obtained from Europe Space Agency (ESA, [25]). Nine different regions of interest are outlined as red rectangles, including Forest in Northern Canada (FnC), Forest in Eastern Russia (FeR), Cropland in Western Russia (CwR), Cropland in Northern China (CnC), Cropland in Northern India (CnI), Cropland in Eastern Brazil (CeB), Grassland in Central Africa (GcA), Mixed forest in Southern Africa (FsA) and Shrub-land in Northern Australia (SnA). Six regions in the Northern Hemisphere, including FnC, CwR, FeR, CnC, CnI, and GcA, are used for correlation analysis of seasonal cycle between NDVI and XCO2 in Section 3.2, while five regions, including CnI, GcA, CeB, FsA, and SnA, are used for correlation analysis of seasonal variations between LST and XCO2 in Section 4.2. Twelve selected sites of TCCON are marked with black dots and used for evaluating mapping data in Section 4.1.2.
Remotesensing 09 00251 g001
Figure 2. Example of fitting GM-XCO2 using Equation (1) in the top panel, in which (k0, k1, a1, a2, b1, b2) = (386.058, 0.0163, −0.555, 1.116, −3.731, −1.311). In the bottom panel, the seasonal GM-XCO2 (dXCO2) is calculated by removing the yearly increases from the temporal variation of GM-XCO2. The fitting residual is the difference between GM-XCO2 and the fitting. Monthly-averaged seasonal GM-XCO2 is the monthly mean of dXCO2, which is used in Section 3.2.
Figure 2. Example of fitting GM-XCO2 using Equation (1) in the top panel, in which (k0, k1, a1, a2, b1, b2) = (386.058, 0.0163, −0.555, 1.116, −3.731, −1.311). In the bottom panel, the seasonal GM-XCO2 (dXCO2) is calculated by removing the yearly increases from the temporal variation of GM-XCO2. The fitting residual is the difference between GM-XCO2 and the fitting. Monthly-averaged seasonal GM-XCO2 is the monthly mean of dXCO2, which is used in Section 3.2.
Remotesensing 09 00251 g002
Figure 3. Spatial distribution of SCA of XCO2 in global land areas derived from the GM-XCO2 dataset from June 2009 to May 2014. The SCA for each grid is calculated by fitting Equation (1) to the 5-year time series of GM-XCO2 and obtaining the annual mean of differences between peaks and troughs after subtracting the linear trend of the yearly increase. Two contour lines for amplitude of 8.0 ppm and 4.0 ppm are outlined with solid and dotted black lines, respectively.
Figure 3. Spatial distribution of SCA of XCO2 in global land areas derived from the GM-XCO2 dataset from June 2009 to May 2014. The SCA for each grid is calculated by fitting Equation (1) to the 5-year time series of GM-XCO2 and obtaining the annual mean of differences between peaks and troughs after subtracting the linear trend of the yearly increase. Two contour lines for amplitude of 8.0 ppm and 4.0 ppm are outlined with solid and dotted black lines, respectively.
Remotesensing 09 00251 g003
Figure 4. Spatial distribution of NDVI SCA, which is calculated for each grid by fitting Equation (1) to the 5-year NDVI data and obtaining the annual-mean differences between peaks and troughs after subtracting the linear trend. The two overlapped contour lines are the XCO2 SCA from GM-XCO2, with 8.0 ppm in solid black line and 4.0 ppm in dotted black line as shown in Figure 3.
Figure 4. Spatial distribution of NDVI SCA, which is calculated for each grid by fitting Equation (1) to the 5-year NDVI data and obtaining the annual-mean differences between peaks and troughs after subtracting the linear trend. The two overlapped contour lines are the XCO2 SCA from GM-XCO2, with 8.0 ppm in solid black line and 4.0 ppm in dotted black line as shown in Figure 3.
Remotesensing 09 00251 g004
Figure 5. Pearson Correlation Coefficients (PCCs) between monthly mean of detrended GM-XCO2 (dXCO2) and parameters for biosphere activities, including (a) NDVI, (b) EVI, (c) LAI, and (d) GPP from MODIS data products in each grid from June 2009 to May 2014. Only grids with significant levels of correlation (p-value < 0.01) are shown. In calculation of the correlation coefficient, all data are averaged by month. The overlapped lines are 8.0 ppm in solid black and 4.0 ppm in dotted black as shown in Figure 3.
Figure 5. Pearson Correlation Coefficients (PCCs) between monthly mean of detrended GM-XCO2 (dXCO2) and parameters for biosphere activities, including (a) NDVI, (b) EVI, (c) LAI, and (d) GPP from MODIS data products in each grid from June 2009 to May 2014. Only grids with significant levels of correlation (p-value < 0.01) are shown. In calculation of the correlation coefficient, all data are averaged by month. The overlapped lines are 8.0 ppm in solid black and 4.0 ppm in dotted black as shown in Figure 3.
Remotesensing 09 00251 g005
Figure 6. Seasonal variations of monthly-averaged seasonal GM-XCO2 (dXCO2, black lines) and NDVI (green lines) from June 2009 to May 2014 in six selected regions of interests as shown in Figure 1 and Figure 5a. Vertical error bars represent one standard deviation from its mean value. The PCC values given in the figure titles, between seasonal variations of GM-XCO2 and NDVI, are averaged values for all grids within the region of interest. The time delay, in terms of lags of months, of the NDVI time series to maximize the negative correlation with XCO2 is 1.6 months for FnC, with a minimum averaged PCC of −0.82. They are 1.0 month and −0.85 for FeR, 1.8 months and −0.83 for CwR, 0.6 months and −0.85 for CnC, 0.8 months and −0.86 for GcA, and 0.3 months and −0.81 for CnI.
Figure 6. Seasonal variations of monthly-averaged seasonal GM-XCO2 (dXCO2, black lines) and NDVI (green lines) from June 2009 to May 2014 in six selected regions of interests as shown in Figure 1 and Figure 5a. Vertical error bars represent one standard deviation from its mean value. The PCC values given in the figure titles, between seasonal variations of GM-XCO2 and NDVI, are averaged values for all grids within the region of interest. The time delay, in terms of lags of months, of the NDVI time series to maximize the negative correlation with XCO2 is 1.6 months for FnC, with a minimum averaged PCC of −0.82. They are 1.0 month and −0.85 for FeR, 1.8 months and −0.83 for CwR, 0.6 months and −0.85 for CnC, 0.8 months and −0.86 for GcA, and 0.3 months and −0.81 for CnI.
Remotesensing 09 00251 g006
Figure 7. (a) The global distribution of phase delay effect, in terms of lags of months, between seasonal variations of XCO2 and NDVI, and (b) the correlation between XCO2 and NDVI if the phase delay effect is excluded. Only grids with month lags of no more than three are shown.
Figure 7. (a) The global distribution of phase delay effect, in terms of lags of months, between seasonal variations of XCO2 and NDVI, and (b) the correlation between XCO2 and NDVI if the phase delay effect is excluded. Only grids with month lags of no more than three are shown.
Remotesensing 09 00251 g007
Figure 8. The relationship between predicted XCO2 and observed XCO2 values in cross-validation of global land mapping of XCO2. The color grids represent the density of data distribution. The dotted line is derived from linear regression of predicted values of XCO2 (Y) and the observed values of XCO2 (X), which shows a significant linear relationship with R2 equals 0.91 (p-value < 0.01) and good consistency of observed XCO2 and predicted XCO2 with MAPE equal to 0.85. The solid line shows the one-to-one line.
Figure 8. The relationship between predicted XCO2 and observed XCO2 values in cross-validation of global land mapping of XCO2. The color grids represent the density of data distribution. The dotted line is derived from linear regression of predicted values of XCO2 (Y) and the observed values of XCO2 (X), which shows a significant linear relationship with R2 equals 0.91 (p-value < 0.01) and good consistency of observed XCO2 and predicted XCO2 with MAPE equal to 0.85. The solid line shows the one-to-one line.
Remotesensing 09 00251 g008
Figure 9. Global distribution of XCO2 SCA derived from CarbonTracker (CT2015) CT-XCO2 dataset from June 2009 to May 2014. Similar to Figure 4, the SCA for each grid is calculated by fitting its XCO2 time series using Equation (1) and obtaining the annual mean of differences between peaks and troughs after subtracting the linear trend. The two overlapped contour lines are from the XCO2 SCA (8.0 ppm in the solid black line and 4.0 ppm in the dotted black line) as shown in Figure 3.
Figure 9. Global distribution of XCO2 SCA derived from CarbonTracker (CT2015) CT-XCO2 dataset from June 2009 to May 2014. Similar to Figure 4, the SCA for each grid is calculated by fitting its XCO2 time series using Equation (1) and obtaining the annual mean of differences between peaks and troughs after subtracting the linear trend. The two overlapped contour lines are from the XCO2 SCA (8.0 ppm in the solid black line and 4.0 ppm in the dotted black line) as shown in Figure 3.
Remotesensing 09 00251 g009
Figure 10. XCO2 from GOSAT-ACOS retrievals (black) and CarbonTracker (red) from June 2009 to May 2014 over mid- and high latitude of North Hemisphere. XCO2 over latitude bands between 50°–60°N in (a) Eurasia and (c) North-America land regions, and between 40°–50°N for the same two regions in (b,d), respectively.
Figure 10. XCO2 from GOSAT-ACOS retrievals (black) and CarbonTracker (red) from June 2009 to May 2014 over mid- and high latitude of North Hemisphere. XCO2 over latitude bands between 50°–60°N in (a) Eurasia and (c) North-America land regions, and between 40°–50°N for the same two regions in (b,d), respectively.
Remotesensing 09 00251 g010
Figure 11. Pearson correlation coefficients (PCCs) between the monthly values of detrended XCO2 from GM-XCO2 and the land surface temperature (LST) from the MODIS data products in each grid from June 2009 to May 2014 between 30°S and 30°N. Only grids with a significant level of correlation (p-value < 0.01) are shown. In calculation of correlation coefficient, all data are averaged by month.
Figure 11. Pearson correlation coefficients (PCCs) between the monthly values of detrended XCO2 from GM-XCO2 and the land surface temperature (LST) from the MODIS data products in each grid from June 2009 to May 2014 between 30°S and 30°N. Only grids with a significant level of correlation (p-value < 0.01) are shown. In calculation of correlation coefficient, all data are averaged by month.
Remotesensing 09 00251 g011
Figure 12. Temporal variation of de-trended GM-XCO2 (dXCO2, black lines) and LST (red lines) from June 2009 to May 2014 in five regions of interests. Vertical bars represent the range of ±1σ.
Figure 12. Temporal variation of de-trended GM-XCO2 (dXCO2, black lines) and LST (red lines) from June 2009 to May 2014 in five regions of interests. Vertical bars represent the range of ±1σ.
Remotesensing 09 00251 g012
Table 1. Specifications of the used satellite observations from GOSAT and MODIS and model simulations from CarbonTracker spanning from 1 June 2009 to 31 May 2014.
Table 1. Specifications of the used satellite observations from GOSAT and MODIS and model simulations from CarbonTracker spanning from 1 June 2009 to 31 May 2014.
DatasetResolution Spatial TemporalDescriptionReference
Global mapping XCO2 (GM-XCO2)1 × 1 degree3 dayGlobal land mapping of XCO2 by spatio-temporal geostatistics approach using GOSAT XCO2 (v 3.5) retrievals from ACOS project team.Zeng et al. [20];
O’Dell et al. [26]
Simulated XCO2 (CT-XCO2)2 × 3 degree3 dayGlobal simulations of XCO2 derived from multi-layer CO2 data simulated by CarbonTracker CT2015Peter et al. [27]
NDVI0.05 degreemonthlyNormalized Difference Vegetation Index (NDVI) from MOD09 reflectance product (MYD13C2)Huete et al. [28]
EVI0.05 degreemonthlyEnhanced Vegetation Index (EVI) from MOD09 reflectance product (MYD13C2)Huete et al. [28]
GPP1 kmmonthlyGross Primary Productivity (GPP) from MODIS data product (MOD17A2)Heinsch et al. [29]
LAI1 km8 daysLeaf Area Index (LAI) from MODIS data product (MCD15A2)Yang et al. [30]
LST0.05 degreemonthlyLand Surface Temperature (LST) from MODIS data product (MOD11C3)Wan et al. [31]
Table 2. Specifications of the selected regions of interest including the location, major vegetation types and its area proportion, and averaged Pearson Correlation Coefficients (PCCs) of seasonal variation between GM-XCO2 and NDVI or LST over each region of interest. Several regions of interests are selected for NDVI and LST, respectively.
Table 2. Specifications of the selected regions of interest including the location, major vegetation types and its area proportion, and averaged Pearson Correlation Coefficients (PCCs) of seasonal variation between GM-XCO2 and NDVI or LST over each region of interest. Several regions of interests are selected for NDVI and LST, respectively.
RegionsLong NameLatitudeLongitudeDescriptionPCC with NDVIPCC with LST
FnCForest in Northern Canada50°–57°N86°–99°WF: 63%−0.61-
FeRForest in Eastern Russia52°–59°N117°–130°EF: 86%−0.68-
CwRCropland in Western Russia50°–57°N33°–46°EC: 70%−0.50-
CnCCropland in Northern China35°–42°N107°–120°EC: 58%
G-S: 18%
−0.80-
CnICropland in Northern India 19°–26°N74°–87°EC: 89%−0.790.78
GcAGrassland in Central Africa 5°–12°N15°–28°EG:36%
F: 30%
−0.800.79
CeBCropland in Eastern Brazil8°–15°S40°–53°WC: 43%
F: 15%
-0.49
FsAMixed forest in Southern Africa10°–17°S20°–33°EF: 64%-0.59
SnAShrub-land in Northern Australia11°–18°S124°–137°EG: 60%-0.47
Notes: F: Forest, C: Cropland, G-S: Grassland and Shrub-land, G: Grassland.
Table 3. Statistics of comparison between GM-XCO2 and TCCON data (smoothed by applying the ACOS-GOSAT averaging kernel), and their SCAs for the following 12 sites. Bias is calculated using GM-XCO2 minus TCCON XCO2 for each coincident data pair and averaged for each site.
Table 3. Statistics of comparison between GM-XCO2 and TCCON data (smoothed by applying the ACOS-GOSAT averaging kernel), and their SCAs for the following 12 sites. Bias is calculated using GM-XCO2 minus TCCON XCO2 for each coincident data pair and averaged for each site.
SitesLocation (Latitude, Longitude)Coincident Data PairsAveraged Biases (ppm)Averaged Absolute Bias (ppm)Standard Deviation (ppm)GM-XCO2 SCA (ppm)TCCON SCA (ppm)SCA Difference (ppm)
Bialystok(53.23, 23.02)2310.331.102.037.388.38−1.00
Bremen(53.1, 8.85)119−0.351.232.676.677.35−0.68
Karlsruhe(49.1, 8.44)193−0.571.332.596.127.50−1.38
Orleans(47.97, 2.11)2590.211.111.945.877.62−1.75
Garmisch(47.48, 11.06)322−0.491.212.095.996.99−1.00
Park Falls(45.94, −90.27)453−0.241.021.629.259.20.05
Lamont(36.6, −97.49)5130.250.951.405.185.8−0.62
Tsukuba(36.05, 140.12)246−1.721.942.707.386.381.00
JPL/Caltech(34.2, −118.18)1950.101.292.455.326.07−0.75
Saga(33.24, 130.29)2081.081.371.396.946.650.29
Darwin(−12.43, 130.89)413−0.361.021.372.271.121.15
Wollongong(−34.41, 150.88)412−0.360.680.651.341.41−0.07
Overall 3564−0.181.191.91 −0.40

Share and Cite

MDPI and ACS Style

He, Z.; Zeng, Z.-C.; Lei, L.; Bie, N.; Yang, S. A Data-Driven Assessment of Biosphere-Atmosphere Interaction Impact on Seasonal Cycle Patterns of XCO2 Using GOSAT and MODIS Observations. Remote Sens. 2017, 9, 251. https://doi.org/10.3390/rs9030251

AMA Style

He Z, Zeng Z-C, Lei L, Bie N, Yang S. A Data-Driven Assessment of Biosphere-Atmosphere Interaction Impact on Seasonal Cycle Patterns of XCO2 Using GOSAT and MODIS Observations. Remote Sensing. 2017; 9(3):251. https://doi.org/10.3390/rs9030251

Chicago/Turabian Style

He, Zhonghua, Zhao-Cheng Zeng, Liping Lei, Nian Bie, and Shaoyuan Yang. 2017. "A Data-Driven Assessment of Biosphere-Atmosphere Interaction Impact on Seasonal Cycle Patterns of XCO2 Using GOSAT and MODIS Observations" Remote Sensing 9, no. 3: 251. https://doi.org/10.3390/rs9030251

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop