This is not the latest CarbonTracker update! Link to latest.
 
Documentation - CT2011_oi
Biosphere Oceans Observations Fires Fossil Fuel TM5 Nested Model Assimilation
To learn more about a CarbonTracker component, click on one of the above images.
Or download the full PDF version for convenience.

Ensemble Data Assimilation [goto top]
1.   Introduction
Data assimilation is the name of a process by which observations of the 'state' of a system help to constrain the behavior of the system in time. An example of one of the earliest applications of data assimilation is the system in which the trajectory of a flying rocket is constantly (and rapidly) adjusted based on information of its current position, heading, speed, and other factors, to guide it to its exact final destination. Another example of data assimilation is a weather model that gets updated every few hours with measurements of temperature and other variables, to improve the accuracy of its forecast for the next day, and the next, and the next. Data assimilation is usually a cyclical process, as estimates get refined over time as more observations about the "truth" become available. Mathematically, data assimilation can be done with any number of techniques. For large systems, so-called variational and ensemble techniques have gained most popularity. Because of the size and complexity of the systems studied in most fields, data assimilation projects inevitably include supercomputers that model the known physics of a system. Success in guiding these models in time often depends strongly on the number of observations available to inform on the true system state.

In CarbonTracker, the model that describes the system contains relatively simple descriptions of biospheric and oceanic CO2 exchange, as well as fossil fuel and fire emissions. In time, we alter the behavior of this model by adjusting a small set of parameters as described in the next section.

2.   Detailed Description
The four surface flux modules drive instantaneous CO2 fluxes in CarbonTracker according to:

F(x, y, t) = λ • Fbio(x, y, t) + λ • Foce(x, y, t) + Fff(x, y, t) + Ffire(x, y, t)

Where λ represents a set of linear scaling factors applied to the fluxes, to be estimated in the assimilation. These scaling factors are the final product of our assimilation and together with the modules determine the fluxes we present in CarbonTracker. Note that no scaling factors are applied to the fossil fuel and fire modules.

2.1   Land-surface classification
The scaling factors λ are estimated for each week and assumed constant over this period. Each scaling factor is associated with a particular region of the global domain, and currently the geographical distribution of the regions is fixed. The choice of regions is a strong a-priori constraint on the resulting fluxes and should be approached with care to avoid so-called "aggregation errors" [Kaminski et al., 2001]. We chose an approach in which the ocean is divided up into 30 large basins encompassing large-scale ocean circulation features, as in the TransCom inversion study (e.g. Gurney et al., [2002]). The terrestrial biosphere is divided up according to ecosystem type as well as geographical location. Thereto, each of the 11 TransCom land regions contains a maximum of 19 ecosystem types summarized in the table below. Figure 1 shows ecoregions for North America (click here for global land ecoregions). Note that there is currently no requirement for ecoregions to be contiguous, and a single scaling factor can be applied to the same vegetation type on both sides of a continent. Further details on ecoregions can be found here.

Fig 1. CarbonTracker ecoregions in North America

Theoretically, this approach leads to a total number of 11*19+30=239 optimizable scaling factors λ each week, but the actual number is 156 since not every ecosystem type is represented in each TransCom region, and because we decided not to optimize parameters for ice-covered regions, inland water bodies, and desert. The total flux coming out of these last regions is negligibly small. It is important to note that even though only one parameter is available to scale, for instance, the flux from coniferous forests in Boreal North America, each 1° x 1° grid box predominantly covered by coniferous forests will have a different flux F(x,y,t) depending on local temperature, radiation, and CASA modeled monthly mean flux.

Ecosystem types considered on 1° x 1° for the terrestrial flux inversions is based on Olson, [1992]. Note that we have adjusted the original 29 categories into only 19 regions. This was done mainly to fill the unused categories 16,17, and 18, and to group the similar (from our perspective) categories 23-26+29. The table below shows each vegetation category considered. Percentages indicate the area associated with each category for North America rounded to one decimal.

Ecosystem Types
categoryOlson V 1.3aPercentage area
1Conifer Forest 19.0%
2Broadleaf Forest 1.3%
3Mixed Forest 7.5%
4Grass/Shrub 12.6%
5Tropical Forest 0.3%
6Scrub/Woods 2.1%
7Semitundra 19.4%
8Fields/Woods/Savanna 4.9%
9Northern Taiga 8.1%
10Forest/Field 6.3%
11Wetland 1.7%
12Deserts 0.1%
13Shrub/Tree/Suc 0.1%
14Crops 9.7%
15Conifer Snowy/Coastal 0.4%
16Wooded tundra 1.7%
17Mangrove 0.0%
18Non-optimized areas (ice, polar desert, inland seas) 0.0%
19Water 4.9%

Each 1° x 1° pixel of our domain was assigned one of the categories above bases on the Olson category that was most prevalent in the 0.5° x 0.5° underlying area.

2.2   Ensemble Size and Localization
The ensemble system used to solve for the scalar multiplication factors is similar to that in Peters et al. [2005] and based on the square root ensemble Kalman filter of Whitaker and Hamill, [2002]. We have restricted the length of the smoother window to only five weeks as we found the derived flux patterns within North America to be robustly resolved well within that time. We caution the CarbonTracker users that although the North American flux results were found to be robust after five weeks, regions of the world with less dense observational coverage (the tropics, Southern Hemisphere, and parts of Asia) are likely to be poorly observable even after more than a month of transport and therefore less robustly resolved. Although longer assimilation windows, or long prior covariance length-scales, could potentially help to constrain larger scale emission totals from such areas, we focus our analysis here on a region more directly constrained by real atmospheric observations.

Ensemble statistics are created from 150 ensemble members, each with its own background CO2 concentration field to represent the time history (and thus covariances) of the filter. To dampen spurious noise due to the approximation of the covariance matrix, we apply localization [Houtekamer and Mitchell, 1998] for non-MBL sites only. This ensures that tall-tower observations within North America do not inform on for instance tropical African fluxes, unless a very robust signal is found. In contrast, MBL sites with a known large footprint and strong capacity to see integrated flux signals are not localized. Localization is based on the linear correlation coefficient between the 150 parameter deviations and 150 observation deviations for each parameter. If the relationship between a parameter deviation and its modeled observational impact is statistically significant, then that relationship is used to modify parameters. Otherwise, the relationship is assumed to be spurious noise due to the numerical approximation of the covariance matrix by the limited ensemble. We accept relationships that reach 95% significance in a student's T-test with a two-tailed probability distribution.

2.3   Dynamical Model
In CarbonTracker, the dynamical model is applied to the mean parameter values λ as:

λ tb = (λ  t-2a + λ  t-1 a + λ  p  )   ⁄   3.0

Where "a" refers to analyzed quantities from previous steps, "b" refers to the background values for the new step, and "p" refers to real a-priori determined values that are fixed in time and chosen as part of the inversion set-up. Physically, this model describes that parameter values λ for a new time step are chosen as a combination between optimized values from the two previous time steps, and a fixed prior value. This operation is similar to the simple persistence forecast used in Peters et al. [2005], but represents a smoothing over three time steps thus dampening variations in the forecast of λ b in time. The inclusion of the prior term λ p acts as a regularization [Baker et al., 2006] and ensures that the parameters in our system will eventually revert back to predetermined prior values when there is no information coming from the observations. Note that our dynamical model equation does not include an error term on the dynamical model, for the simple reason that we don't know the error of this model. This is reflected in the treatment of covariance, which is always set to a prior covariance structure and not forecast with our dynamical model.

3   Covariance Structure
Prior values for λ p are all 1.0 to yield fluxes that are unchanged from their values predicted in our modules. The prior covariance structure Pp describes the magnitude of the uncertainty on each parameter, plus their correlation in space. The latter is applied such that correlations between the same ecosystem types in different TransCom regions decrease exponentially with distance (L=2000km), and thus assumes a coupling between the behavior of the same ecosystems in close proximity to one another (such as coniferous forests in Boreal and Temperate North America). Furthermore, all ecosystems within tropical TransCom regions are coupled decreasing exponentially with distance since we do not believe the current observing network can constrain tropical fluxes on sub-continental scales, and want to prevent spurious compensating source/sink pairs ("dipoles") to occur in the tropics.

In our standard assimilation, the chosen standard deviation is 80% on land parameters. All parameters have the same variance within the land or ocean domain. Because the parameters multiply the net-flux though, ecosystems with larger weekly mean net fluxes have a larger variance in absolute flux magnitude.

3.1   Multiple prior models
In Bayesian estimation systems like CarbonTracker, there is a potential for bias from a flux prior to propagate through the inversion system to the final result. It is difficult to quantify this effect, and as a result it is generally considered a precondition that flux priors be unbiased. We cannot guarantee this for any of our fluxes, be they the prior estimates for terrestrial or oceanic exchange, or the presumed wildfire and fossil fuel emissions. In order to explicitly quantify the impact of prior bias on our solution, in CT2011 CT2011_oi we present the result of a multi-model prior suite of inversions. We have used two terrestrial flux priors, two air-sea exchange priors one air-sea exchange prior, and two estimates of imposed fossil fuel emissions in a three-way factorial design experiment. This has resulted in eight four individual inversions, each using a unique combination of priors and conducted independently according to the methods described above. We present as a final result the mean flux across this suite of inversions and the atmospheric CO2 distribution resulting from applying these mean fluxes to our atmospheric transport model. Each of the priors is described in detail on its corresponding documentation page (fossil, land, ocean).

Notice: CT2011_oi does not use the "climatological" ocean prior. After we released CarbonTracker 2011, a significant bug was discovered in our atmospheric transport model. We have corrected the bug and are releasing revised results under the release name "CT2011_oi". One consequence of this problem is that the four inversions using the climatological ocean flux prior were faulty. They have been removed from the inversion suite in CT2011_oi. Use of the original CT2011 results is strongly discouraged. Details can be found at this link.

Fig 2. CT2011_oi prior covariance structure. The prior covariance matrix (top panel) and the square root of diagonal members of this matrix (bottom panel). Covariance matrix quantities are dimensionless squared scaling factors, and the bottom panel is the square root of this. TransCom land regions form the first 11 large divisions on the axes here. As described above, each of those regions contains 19 potential ecosystems. Correlations between similar ecosystems in proximate Transcom regions are visible in North America (e.g. NABR and NATM, the boreal and temperate North American regions) and Eurasia. Within tropical Transcom regions, however, differing ecosystems are assigned a non-zero prior covariance, which is visible here as red block-like structures on the diagonal within, for example, the South America Tropical (SATR) Transcom region. Ocean regions have a more complicated covariance structure that depends on which prior is used; the structure shown here is that of the ocean inversion flux prior. The lower panel of this diagram compares the on-diagonal elements of the prior covariance matrix by plotting their square roots. The resulting standard deviations are directly comparable to the percentages discussed in section 3 above; 0.8 is equivalent to 80%. The retuning of the covariance matrix for CT2011_oi's multiple-prior simulation is made evident by also showing these values from previous CarbonTracker releases in red.


3.2   Posterior Uncertainties in CarbonTracker

The formal "internal" error estimates produced by CarbonTracker are unrealistically large. This is largely a result of the relatively short assimilation window in CarbonTracker, along with a dynamical model that introduces a fresh prior covariance matrix with every new week entering the assimilation window. This five-week window effectively inhibits the formation of anticorrelations ("dipoles") in flux estimates, and does little to reduce the confidence interval on prior fluxes.

The temporal truncation in CarbonTracker imposed by its five-week assimilation window tends to yield regional flux estimates that are largely uncorrelated with those from other regions. A consequence of this feature is that uncertainties in CarbonTracker tend to increase as larger regions are considered; regional errors mostly just add in quadrature without any cancellation from dipole anticorrelation. Whereas many inversions yield smaller errors as the spatial extent of the region being considered increases, CarbonTracker acts in the opposite fashion. This is perhaps most obvious in the estimate of CarbonTracker's global annual surface flux of carbon dioxide. While CT2011_oi estimates a one-sigma error of more than 6 PgCyr-1 on its global flux, this quantity is in actuality much more well-constrained. This is evident from CarbonTracker's excellent agreement with observational estimates of atmospheric growth rate.

In CT2011_oi, error estimates are about a factor of two larger than in previous releases, mainly due to the retuning of the land prior covariance discussed above. However, uncertainties presented for CT2011_oi take into account not only the "internal" flux uncertainty generated by a single inversion, but also the across-model "external" uncertainty representing the spread of the inversion models due to the choice of prior flux.

4.   Further Reading