CarbonTracker Documentation
CT2022 release

Originally published Feb 17, 2023
Revised, Mar 3, 2023

Andrew R. Jacobson1,2, Kenneth N. Schuldt1,2, Pieter Tans3, Arlyn Andrews2,
John B. Miller2, Tomohiro Oda4,5,6, Sourish Basu7,8, John Mund1,2,
Brad Weir9,7, Lesley Ott7, Tuula Aalto10, James Brice Abshire11,
Ken Aikin12, Shuji Aoki13, Francesco Apadula14, Sabrina Arnold15,
Bianca Baier2, Jakub Bartyzel16, Andreas Beyersdorf17, Tobias Biermann18,
Sebastien C. Biraud19, Harald Boenisch20, Gordon Brailsford21, Willi A. Brand22,
Gao Chen23, Huilin Chen24, Lukasz Chmura16, Shane Clark25,
Aurelie Colomb26, Roisin Commane27, Sébastien Conil28, Cédric Couret29,
Adam Cox25, Paolo Cristofanelli30, Emilio Cuevas31, Roger Curcoll32,
Bruce Daube27, Kenneth J. Davis33, Stephan De Wekker34, Julian Della Coletta35,
Marc Delmotte36, Elizabeth DiGangi37, Joshua P. DiGangi23, Alcide Giorgio di Sarra38,
Ed Dlugokencky2, James W. Elkins2, Lukas Emmenegger39, Shuangxi Fang40,
Marc L. Fischer41, Grant Forster42,43, Arnoud Frumau44, Michal Galkowski16,
Luciana V. Gatti45, Torsten Gehrlein20, Christoph Gerbig22, Francois Gheusi46,
Emanuel Gloor47, Vanessa Gomez-Trueba48,31, Daisuke Goto49, Tim Griffis50,
Samuel Hammer35, Chad Hanson51, László Haszpra52, Juha Hatakka10,
Martin Heimann22, Michal Heliasz18, Arjan Hensen44, Ove Hermansen53,
Eric Hintsa2, Jutta Holst54, Viktor Ivakhov55, Daniel A. Jaffe56,
Armin Jordan22, Warren Joubert57, Anna Karion58, Stephan Randolph Kawa11,
Victor Kazan36, Ralph F. Keeling25, Petri Keronen59, Tobias Kneuer60,
Pasi Kolari59, Kateřina Komínková61, Eric Kort62, Elena Kozlova63,
Paul Krummel64, Dagmar Kubistin60, Casper Labuschagne57, David H.Y. Lam65,
Xin Lan1,2, Ray L. Langenfelds64, Olivier Laurent66, Tuomas Laurila10,
Thomas Lauvaux67,33, Jost Lavric22, Beverly E. Law51, John Lee68,
Olivia S.M. Lee65, Irene Lehner18, Kari Lehtinen10, Reimo Leppert22,
Ari Leskinen69,70, Markus Leuenberger71, Ingeborg Levin35, Janne Levula59,
John Lin72, Matthias Lindauer60, Zoe Loh64, Morgan Lopez36,
Ingrid T. Luijkx73,74, Chris René Lunder53, Toshinobu Machida75, Ivan Mammarella76,
Giovanni Manca77, Alistair Manning78, Andrew Manning43, Michal V. Marek61,
Melissa Yang Martin23, Hidekazu Matsueda79, Kathryn McKain2, Harro Meijer24,
Frank Meinhardt80, Lynne Merchant25, N. Mihalopoulos81, Natasha L. Miles33,
Charles E. Miller82, Logan Mitchell72, Meelis Mölder54, Stephen Montzka2,
Fred Moore2, Heiko Moossen22, Eric Morgan25, Josep-Anton Morgui32,
Shinji Morimoto13, Jennifer Müller-Williams60, J. William Munger27, David Munro2,
Cathrine Lund Myhre53, Shin-Ichiro Nakaoka75, Jaroslaw Necki16, Sally Newman83,
Sylvia Nichol21, Yosuke Niwa75, Florian Obersteiner20, Simon O'Doherty84,
Bill Paplawsky25, Jeff Peischl1,12, Olli Peltola59, Salvatore Piacentino38,
Jean-Marc Pichon26, Penelope Pickers43, Steve Piper25, Joseph Pitt84,
Christian Plass-Dülmer60, Stephen Matthew Platt53, Steve Prinzivalli37, Michel Ramonet36,
Ramon Ramos31, Enrique Reyes-Sanchez31, Scott J. Richardson33, Haris Riris11,
Pedro P. Rivas31, Thomas Ryerson12, Kazuyuki Saito85, Maryann Sargent27,
Motoki Sasakawa75, Bert Scheeren24, Tanja Schuck86, Marcus Schumacher22,
Thomas Seifert22, Mahesh Kumar Sha87, Paul Shepson88, Michael Shook23,
Christopher D. Sloop37, Paul Smith89, Kieran Stanley84, Martin Steinbacher39,
Britton Stephens90, Colm Sweeney2, Kirk Thoning2, Helder Timas91,
Margaret Torn41, Kjetil Tørseth53, Pamela Trisolino30, Jocelyn Turnbull92,1,
Pim van den Bulk44, Danielle van Dinther44, Alex Vermeulen54, Brian Viner93,
Gabriela Vitkova61, Stephen Walker25, Andrew Watson63, Steven C. Wofsy27,
Justin Worsey63, Doug Worthy94, Dickon Young84, Sönke Zaehle22,
Andreas Zahn20 and Miroslaw Zimnoch16

1CIRES, University of Colorado, Boulder, Colorado, USA
2NOAA Global Monitoring Laboratory, Boulder, Colorado, USA
3Institute of Arctic and Alpine Research,University of Colorado, Boulder, Colorado, USA
4Earth from Space Institute, Universities Space Research Association, Washington, DC, USA
5Department of Atmospheric and Oceanic Science, University of Maryland, College Park, Maryland, USA
6Graduate School of Engineering, Osaka University, Suita, Osaka, Japan
7Global Modeling and Assimilation Office, NASA Goddard Space Flight Center, Greenbelt, Maryland, USA
8Earth System Science Interdisciplinary Center, University of Maryland, College Park, Maryland, USA
9Morgan State University, Baltimore, Maryland, USA
10Finnish Meteorological Institute, Climate System Research, Helsinki, Finland
11NASA Goddard Space Flight Center, Greenbelt, Maryland, USA
12NOAA Chemical Sciences Laboratory, Boulder, Colorado, USA
13Tohoku University, Sendai, Japan
14Ricerca sul Sistema Energetico– RSE S.p.A., Milano, Italy
15Deutsches Zentrum für Luft- und Raumfahrt (DLR), Institut für Physik der Atmosphäre, Oberpfaffenhofen, Germany
16AGH University of Science and Technology, Krakow, Poland
17California State University, San Bernardino, California, USA
18Lund University, Centre for Environmental and Climate Science, Lund, Sweden
19ARM Carbon Project, Lawrence Berkeley National Laboratory, Berkeley, California, USA
20Institute for Meteorology and Climate Research (IMK), Karlsruhe Institute of Technology (KIT), Karlsruhe, Germany
21National Institute of Water and Atmospheric Research, Wellington, New Zealand
22Max Planck Institute for Biogeochemistry, Jena, Germany
23NASA Langley Research Center, Hampton, Virginia, USA
24Centre for Isotope Research, University of Groningen, Groningen, Netherlands
25Scripps Institution of Oceanography, University of California, La Jolla, California, USA
26Observatoire de Physique du Globe de Clermont Ferrand, Aubiere, France
27Harvard University, School of Engineering and Applied Sciences, Cambridge, Massachusetts, USA
28Agence Nationale pour la Gestion des Déchets Radioactifs, France
29Umweltbundesamt, Zugspitze, Germany
30Institute of Atmospheric Sciences and Climate (CNR-ISAC), Bologna, Italy
31Agencia Estatal Meteorologia, Santa Cruz de Tenerife, Spain
32Institut de Ciencia i Tecnologia Ambientals, Universitat Autonoma de Barcelona, Barcelona, Spain
33The Pennsylvania State University, Department of Meteorology and Atmospheric Science, University Park, Pennsylvania, USA
34University of Virginia, Charlottesville, Virginia, USA
35Universität Heidelberg, Institut für Umweltphysik, Heidelberg, Germany
36Laboratoire des Sciences du Climat et de l'Environnement, LSCE/IPSL, CEA-CNRS-UVSQ, Université Paris-Saclay, Gif-sur-Yvette, France
37Earth Networks, Inc., an AEM company, Germantown, Maryland, USA
38Italian National Agency for New Technologies, Energy and Sustainable Economic Development, UTMEA-TER Earth Observations and Analyses Laboratory, Rome, Italy
39Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory for Air Pollution/Environmental Technology, Dübendorf, Switzerland
40Meteorological Observation Centre, Chinese Meteorological Administration, Beijing, China
41Lawrence Berkeley National Laboratory, Berkeley, California, USA
42National Centre for Atmospheric Sciences, University of East Anglia, Norwich, Norfolk, United Kingdom
43Centre for Ocean and Atmospheric Sciences, University of East Anglia, Norfolk, United Kingdom
44Netherlands Organisation for Applied Scientific Research (TNO), Petten, The Netherlands
45National Institute for Space Research (INPE), Sao Paulo, Brazil
46Observatoire Midi-Pyrénées, Toulouse, France
47University of Leeds,School of Geography, Leeds, United Kingdom
48Air Liquide España, Madrid, Spain
49National Institute of Polar Research, Tokyo, Japan
50University of Minnesota,Department of Soil, Water, and Climate, St. Paul, Minnesota, USA
51Oregon State University, Corvallis, Oregon, USA
52Institute for Nuclear Research, Debrecen, Hungary
53NILU, Kjeller, Norway
54Lund University, Dept. Phys. Geography and Ecosystem Science, Lund, Sweden
55Voeikov Main Geophysical Observatory,Saint Petersburg, Russia
56University of Washington, Seattle, Washington, USA
57South African Weather Service, Cape Point, South Africa
58National Institute of Standards and Technology, Gaithersburg, Maryland, USA
59University of Helsinki, Helsinki, Finland
60Deutscher Wetterdienst, Hohenpeißenberg Meteorological Observatory, Hohenpeißenberg, Germany
61Global Change Research Institute of the Czech Academy of Sciences, Brno, Czech Republic
62University of Michigan, Ann Arbor, Michigan, USA
63University of Exeter, Centre for Environmental Data Analysis, Exeter, Devon, United Kingdom
64Commonwealth Scientific and Industrial Research Organisation, Environment, Aspendale, Victoria, Australia
65Hong Kong Observatory, Hong Kong, China
66ICOS Atmospheric Thematic Centre, Gif-sur-Yvette, France
67University of Reims Champagne-Ardenne, CNRS, Reims, France
68University of Maine, Orono, Maine, USA
69University of Eastern Finland, Department of Technical Physics, Kuopio, Finland
70Finnish Meteorological Institute, Kuopio, Finland
71Climate and Environmental Physics, University of Bern, Bern, Switzerland
72Department of Atmospheric Sciences, University of Utah, Salt Lake City, Utah, USA
73Wageningen University, Wageningen, Netherlands
74 ICOS Carbon Portal, Lund University, Lund, Sweden
75National Instiute for Environmental Studies, Tsukuba, Japan
76Institute for Atmospheric and Earth System Research/Physics, Faculty of Sciences, University of Helsinki, Finland
77European Commission, Joint Research Centre, Ispra, Italy
78Met Office Exeter, Devon, United Kingdom
79Meteorological Research Institute, Tsukuba, Japan
80Umweltbundesamt, Oberried-Hofsgrund, Germany
81Environmental and Chemical Processes Laboratory, University of Crete, Crete, Greece
82Jet Propulsion Laboratory, California Institute of Technology, Pasadena California, USA
83California Institute of Technology, Pasadena, California, USA
84University of Bristol, Bristol, United Kingdom
85Japan Meteorological Agency, Tokyo, Japan
86Institute for Atmospheric and Environmental Sciences, University of Frankfurt, Frankfurt, Germany
87Royal Belgian Institute for Space Aeronomy, Brussels, Belgium
88Purdue University, West Lafayette, Indiana, USA
89Svartberget Field Research Station, Swedish University of Agricultural Sciences, Vindeln, Sweden
90National Center for Atmospheric Research, Boulder, Colorado, USA
91Instituto Nacional de Meteorologia e Geofisica, Cidade de Espargos, Ilha do Sal, República de Cabo Verde
92GNS Science,National Isotope Centre, Lower Hutt, New Zealand
93Savannah River National Laboratory, Aiken, South Carolina, USA
94Environment and Climate Change Canado, Ontario, Canada

Contents

1  Introduction
    1.1  A tool for science, and for policy
    1.2  A community effort
    1.3  The role of other atmospheric species in constraining the atmospheric carbon budget
    1.4  Updates
    1.5  Citation and usage policy
        1.5.1  Usage Policy
        1.5.2  Citing our results
2  Terrestrial biosphere module
    2.1  CASA model
    2.2  Temporal downscaling
        2.2.1  Smooth month-to-month variations
    2.3  GFED4.1s and GFED_CMS
3  Fire module
    3.1  Global Fire Emissions Database (GFED)
    3.2  GFED_CMS: Fluxes from the NASA Carbon Monitoring System
4  Fossil fuel module
    4.1  The "Miller" emissions dataset
    4.2  The "ODIAC" emissions dataset
    4.3  Uncertainties
5  Oceans module
    5.1  Air-sea gas exchange
    5.2  OIF: the Ocean Inversion Fluxes prior
    5.3  pCO2-Clim: Takahashi climatology prior
    5.4  Gas-transfer velocity and ocean surface properties
    5.5  Specifics of the inversion methodology related to air-sea CO2 fluxes
6  Atmospheric transport
    6.1  TM5 offline tracer transport model
    6.2  Convective flux fix
7  Observations
    7.1  The CarbonTracker observational network
    7.2  Adaptive model-data mismatch
    7.3  Statistical performance of CT2022
8  Ensemble data assimilation
    8.1  Parameterization of unknowns
        8.1.1  Optimization regions
        8.1.2  Assimilation window
        8.1.3  Ensemble size and localization
        8.1.4  Dynamical model
    8.2  Covariance structure
    8.3  Multiple prior models
        8.3.1  Posterior uncertainties in CarbonTracker
9  Ecoregions in CarbonTracker
    9.1  What are ecoregions?
    9.2  Why use ecoregions?
    9.3  Ecosystems within Transcom regions
10  Resources and References

1  Introduction

The goal of the CarbonTracker program is to produce quantitative estimates of atmospheric carbon uptake and release at the Earth's surface that are consistent with observed patterns of CO2 in the atmosphere. CarbonTracker is an inverse model of atmospheric CO2, which means that it attempts to model atmospheric CO2 measurements by adjusting inputs and removals of CO2 at the Earth surface until they best agree with those observational constraints. CarbonTracker is updated on a approximately-annual basis. The current release, CT2022, provides results from 2000 through the end of 2020. A "near-real" time model product, CT-NRT, extends these results through 2022 and later.

1.1  A tool for science, and for policy

CarbonTracker is made possible by the long-term monitoring of atmospheric CO2 conducted by many academic and governmental programs around the world (see Sec. 7). These data help improve our understanding of how the land and ocean are responding to Earth's changing climate. The uptake and release of CO2 by these ecosystems is changing due to chemical and physical responses to increased atmospheric CO2 concentrations, to human management of lands and oceans, and to changes in temperature, precipitation, and winds.
CarbonTracker is a completely open product. All results, including graphics and tabular data, may be freely used without restriction, although we do request the favor of appropriate acknowledgment (see Sec. 1.5 and https://gml.noaa.gov/ccgg/carbontracker/citation.php).
The unrestricted access to all CarbonTracker results means that anyone can scrutinize our work, suggest improvements, and profit from our efforts. We hope this scrutiny will help guide further development of our methods, and improve our ability to monitor, diagnose, and possibly predict the behavior of the global carbon cycle.
CarbonTracker also can be relevant for helping to inform carbon policy. Its ability to accurately quantify natural and anthropogenic emissions and uptake at regional scales is currently limited by a sparse observational network. With enough observations however, CarbonTracker and systems like it will be able to monitor regional emissions, including those from fossil fuel use. This will provide an independent check on emissions accounting, including estimates of fossil fuel use based on economic inventories. It can thus provide feedback to policies aimed at limiting greenhouse gas emissions. This independent evaluation of the effectiveness of carbon policy is the bottom line in any mitigation strategy. It has the added advantage of being a constraint provided by the atmosphere itself, where CO2 levels matter most.

1.2  A community effort

CarbonTracker is intended to be a tool for the community, and we welcome feedback and collaboration from anyone interested. Our ability to accurately track carbon with more spatial and temporal detail is fundamentally dependent on our collective ability to make enough measurements to characterize variability present in the atmosphere. For example, estimates suggest that observations from tall communication towers (taller than 200m) can tell us about carbon uptake and emission over a radius of only several hundred kilometers. The map of observation sites (Fig. 14) shows how sparse the current network is. One way to join this effort is by contributing measurements to the GLOBALVIEW+ project. Regular air samples collected from the surface, towers or aircraft are needed. It would also be very fruitful to expand use of continuous measurements like the ones now being made on very tall (more than 200m) communications towers. Another way to join this effort is by volunteering flux estimates from your own work, to be run through CarbonTracker and assessed against atmospheric CO2 measurements. We also encourage collaborations focused on use of the CarbonTracker model as a tool for scientific analysis. Please contact us if you would like to get involved and collaborate with us.

1.3  The role of other atmospheric species in constraining the atmospheric carbon budget

Many laboratories making high accuracy CO2 observations also make many other measurements of the same air, typically other greenhouse gases such as methane (CH4), nitrous oxide (N2O), sulfur hexafluoride (SF6), as well as carbon monoxide (CO) and isotopic ratios of CO2 and CH4. These measurements are usually made as mole fractions, for reasons explained here.
These trace gases are relevant for the study of climate change and interesting in their own right, but the additional measurements can also help in identifying sources and sinks of carbon or in understanding carbon cycle processes. For this reason, many air samples are now analyzed for a suite of halocompounds and hydrocarbons. Several of these species can be useful for monitoring air quality, but they can also help with better source apportionment of the greenhouse gases. In addition, the estimation of the source strengths of a number of pollutants could be greatly improved if we were able to quantify fossil fuel CO2 emissions from air measurements for specified regions.
The best tracer for quantifying the component of atmospheric CO2 that has been recently added to an air mass through the burning of fossil fuels is the decrease of the carbon-14 (14C) content of CO2. Cosmic rays produce 14C, a radioactive form of carbon, in the higher regions of the atmosphere. It is present in the atmosphere and oceans and in all living organisms and their remains, but coal, oil, and natural gas contain no 14C because it has long decayed away. Currently, 14CO2 measurements are made on only a small subset of the air samples because of higher analysis costs. None of these other data and their relationships have been used directly in this release of CarbonTracker. We expect them to be incorporated gradually at later stages.
CarbonTracker is a NOAA contribution to the North American Carbon Program.

1.4  Updates

CarbonTracker is updated about once per year to include new data and model improvements. CT2022 provides results from 2000 through 2020. Previous versions of CarbonTracker and our CT-NRT (CarbonTracker Near-Real Time) releases are available at the CarbonTracker website.
Important revisions of our methods for CT2022 include the following:
  • Use of ERA5 reanalysis, including increased vertical resolution,
  • Extension through the end of 2020,
  • Revision of fossil fuel emissions to include COVID pandemic effect,
  • New land and wildfire priors,
  • Cross-evaluation using withheld assimilation data, and
  • Revised model-data mismatch errors on GLOBALVIEW+ measurements

1.5  Citation and usage policy

1.5.1  Usage Policy

CarbonTracker is an open product of NOAA's Global Monitoring Laboratory using data from the NOAA Global Modeling Laboratory greenhouse gas observational network and collaborating institutions. Results, including figures and tabular material found on the CarbonTracker website may be used for non-commercial purposes without restriction. We kindly ask you to acknowledge, cite, and/or reference CarbonTracker as described below.

1.5.2  Citing our results

  • We ask that scientific work that relies heavily on CarbonTracker products is discussed with us before publication, to ensure proper representation of our work and co-authorship if appropriate.
  • Please cite as Jacobson et al. (2023)
  • The DOI for CT2022 and all its associated products and results is http://dx.doi.org/10.25925/z1gj-3254.
  • Please use "CT2022" as the shorthand to refer to our product, not "CT". This identifies both the product and the release version. It is vital to identify the version of the product you are using.
  • Note that the product is called "CarbonTracker" without a space character, not "Carbon ␣ Tracker".
  • Please include our suggested acknowledgment text in your acknowledgments section.
  • Boilerplate model description text provided upon request.
Example  
...we compare our results to NOAA's CarbonTracker, version CT2022 (Jacobson et al., 2023). In this work, CT2022 is ...
Acknowledgments  
CarbonTracker CT2022 results provided by NOAA GML, Boulder, Colorado, USA from the website at http://carbontracker.noaa.gov.
Reference  
Andrew R. Jacobson, Kenneth N. Schuldt, Pieter Tans, Arlyn Andrews, John B. Miller, Tomohiro Oda, Sourish Basu, John Mund, Brad Weir, Lesley Ott, Tuula Aalto, James Brice Abshire, Ken Aikin, Shuji Aoki, Francesco Apadula, Sabrina Arnold, Bianca Baier, Jakub Bartyzel, Andreas Beyersdorf, Tobias Biermann, Sebastien C. Biraud, Harald Boenisch, Gordon Brailsford, Willi A. Brand, Gao Chen, Huilin Chen, Lukasz Chmura, Shane Clark, Aurelie Colomb, Roisin Commane, Sébastien Conil, Cédric Couret, Adam Cox, Paolo Cristofanelli, Emilio Cuevas, Roger Curcoll, Bruce Daube, Kenneth J. Davis, Stephan De Wekker, Julian Della Coletta, Marc Delmotte, Elizabeth DiGangi, Joshua P. DiGangi, Alcide Giorgio di Sarra, Ed Dlugokencky, James W. Elkins, Lukas Emmenegger, Shuangxi Fang, Marc L. Fischer, Grant Forster, Arnoud Frumau, Michal Galkowski, Luciana V. Gatti, Torsten Gehrlein, Christoph Gerbig, Francois Gheusi, Emanuel Gloor, Vanessa Gomez-Trueba, Daisuke Goto, Tim Griffis, Samuel Hammer, Chad Hanson, László Haszpra, Juha Hatakka, Martin Heimann, Michal Heliasz, Arjan Hensen, Ove Hermansen, Eric Hintsa, Jutta Holst, Viktor Ivakhov, Daniel A. Jaffe, Armin Jordan, Warren Joubert, Anna Karion, Stephan Randolph Kawa, Victor Kazan, Ralph F. Keeling, Petri Keronen, Tobias Kneuer, Pasi Kolari, Kateřina Komínková, Eric Kort, Elena Kozlova, Paul Krummel, Dagmar Kubistin, Casper Labuschagne, David H.Y. Lam, Xin Lan, Ray L. Langenfelds, Olivier Laurent, Tuomas Laurila, Thomas Lauvaux, Jost Lavric, Beverly E. Law, John Lee, Olivia S.M. Lee, Irene Lehner, Kari Lehtinen, Reimo Leppert, Ari Leskinen, Markus Leuenberger, Ingeborg Levin, Janne Levula, John Lin, Matthias Lindauer, Zoe Loh, Morgan Lopez, Ingrid T. Luijkx, Chris René Lunder, Toshinobu Machida, Ivan Mammarella, Giovanni Manca, Alistair Manning, Andrew Manning, Michal V. Marek, Melissa Yang Martin, Hidekazu Matsueda, Kathryn McKain, Harro Meijer, Frank Meinhardt, Lynne Merchant, N. Mihalopoulos, Natasha L. Miles, Charles E. Miller, Logan Mitchell, Meelis Mölder, Stephen Montzka, Fred Moore, Heiko Moossen, Eric Morgan, Josep-Anton Morgui, Shinji Morimoto, Jennifer Müller-Williams, J. William Munger, David Munro, Cathrine Lund Myhre, Shin-Ichiro Nakaoka, Jaroslaw Necki, Sally Newman, Sylvia Nichol, Yosuke Niwa, Florian Obersteiner, Simon O\'Doherty, Bill Paplawsky, Jeff Peischl, Olli Peltola, Salvatore Piacentino, Jean-Marc Pichon, Penelope Pickers, Steve Piper, Joseph Pitt, Christian Plass-Dülmer, Stephen Matthew Platt, Steve Prinzivalli, Michel Ramonet, Ramon Ramos, Enrique Reyes-Sanchez, Scott J. Richardson, Haris Riris, Pedro P. Rivas, Thomas Ryerson, Kazuyuki Saito, Maryann Sargent, Motoki Sasakawa, Bert Scheeren, Tanja Schuck, Marcus Schumacher, Thomas Seifert, Mahesh Kumar Sha, Paul Shepson, Michael Shook, Christopher D. Sloop, Paul Smith, Kieran Stanley, Martin Steinbacher, Britton Stephens, Colm Sweeney, Kirk Thoning, Helder Timas, Margaret Torn, Kjetil Tørseth, Pamela Trisolino, Jocelyn Turnbull, Pim van den Bulk, Danielle van Dinther, Alex Vermeulen, Brian Viner, Gabriela Vitkova, Stephen Walker, Andrew Watson, Steven C. Wofsy, Justin Worsey, Doug Worthy, Dickon Young, Sönke Zaehle, Andreas Zahn, Miroslaw Zimnoch. CarbonTracker CT2022, 2023.DOI: 10.25925/z1gj-3254
Acknowledgment text   "CarbonTracker CT2022 results provided by NOAA GML, Boulder, Colorado, USA from the website at http://carbontracker.noaa.gov."
Suggested Website Citation  
"CarbonTracker CT2022, http://carbontracker.noaa.gov"

2  Terrestrial biosphere module

The biospheric component of the terrestrial carbon cycle consists of all the carbon stored in `biomass' around us. This includes trees, shrubs, grasses, carbon within soils, dead wood, and leaf litter. Such reservoirs of carbon can exchange CO2 with the atmosphere. Exchange starts when plants take up CO2 during their growing season through the process of photosynthesis. Most of this carbon is released back to the atmosphere throughout the year through the process of respiration. This process includes both the decay of dead wood and litter and the metabolic respiration of living plants. Of course, plants can also return carbon to the atmosphere when they burn, as described in Section 3. Even though the yearly sum of uptake and release of carbon amounts to a relatively small number, a few petagrams (one Pg=1015 g)) of carbon per year), the flow of carbon each way is as large as 120 Pg C each year. This is why the net result of these flows needs to be monitored in a system such as ours. It is also the reason we need a good physical description-a model-of these flows of carbon. After all, from the atmospheric measurements of CO2 we can see only the relatively small net sum of the much larger two-way streams, or gross fluxes. Information on what the biospheric fluxes are doing in each season, and in every location on Earth is derived from specialized biosphere models, and fed into our system as a first guess, to be refined by our assimilation procedure.

2.1  CASA model

Two biosphere models currently provide first-guess terrestrial fluxes for CT2022. Both models are versions of the Carnegie-Ames Stanford Approach (CASA) biogeochemical model introduced by Potter et al. (1993). CASA calculates global carbon fluxes using input from weather models to drive biophysical processes, and satellite observed Normalized Difference Vegetation Index (NDVI) to track plant phenology. The models are driven by year-specific weather and satellite observations, and include the effects of fires on photosynthesis and respiration (van der Werf et al., 2003), (van der Werf et al., 2006), (Giglio et al., 2006). Both simulations provide 0.5°×0.5° global fluxes with a monthly time resolution.
CASA models directly simulate monthly-mean Net Primary Production (NPP) and heteotrophic respiration (RH) for each terrestrial grid cell being simulated. NPP is the difference in photosynthetic carbon uptake (Gross Primary Production, GPP) and the carbon release by the same plants due to "maintenance respiration", which is also called autotrophic respiration, RA. The carbon uptake represented by NPP and carbon release represented by RH can be differenced to provide Net Ecosystem Exchange (NEE) of CO2. Throughout this discussion, we use the convention that fluxes carry algebraic signs and we adopt the "atmospheric perspective" for those signs. Thus carbon uptake by the terrestrial biosphere is a negative flux to the atmosphere, and release of CO2 back to the atmosphere is a positive flux. This means that we represent all respiration fluxes as positive and GPP as negative, so NEE = NPP + RH. This stands in contrast to convention in the terrestrial carbon community, where all fluxes are generally non-negative.

2.2  Temporal downscaling

Use of monthly-mean terrestrial fluxes to simulate atmospheric CO2 is not sufficient to resolve the variability observed at measurement sites. Instead, higher-frequency variations, including the diurnal cycle and effects of passing weather systems must be imposed on the CASA monthly fluxes. Following the logic laid out by Olsen and Randerson (2004), we transform the CASA-supplied monthly-mean NPP and RH fluxes into GPP and total ecosystem respiration, RE = RA + RH.
To estimate sub-monthly variations, including diurnal and synoptic variability, the Olsen and Randerson (2004) strategy is to model GPP as a linear function of incoming surface solar radiation and total ecosystem respiration as a function of near-surface temperature.
The fundamental assumption needed to apply this scheme is that we can resolve CASA-simulated NPP into GPP and RA. We apply the assumption that GPP is twice NPP, which further implies that RA is the same size as NPP (but of opposite sign):

GPP = 2*NPP,
(1)

NPP = GPP + RA,
(2)
and

RA = −1*NPP.
(3)
We use meteorological fields from the European Centre for Medium-Range Weather Forecasts (ECMWF) ERA5 reanalysis to supply temperature and shortwave radiation. Fluxes are generated with 90-minute variability using a simple temperature Q10 relationship for respiration, assuming a global Q10 value of 1.5, and a linear scaling of photosynthesis with solar radiation. The procedure is very similar, but NOT identical to the procedure in Olsen and Randerson (2004). Note that the introduction of 90-minute variability conserves the monthly mean NEE from the CASA model. Instantaneous NEE for each 90-minute interval is created as:

NEE(t) = GPP(t) + RE(t),
(4)
where

GPP(t) = GPPmean ( I(t) / Imean )
(5)

RE(t) = RE,mean ( Q10(t) / Q10,mean ),
(6)
and Q10 is computed as

Q10(t) = 1.5(T2m(t)−273.15)/10.0 ,
(7)
where T2m is temperature at 2 meters above the land surface in Kelvin, I is surface incoming solar radiation, t is time in 90-minute intervals, and xmean represents the monthly mean of quantity x, including the monthly-mean fluxes derived from the CASA model.

2.2.1  Smooth month-to-month variations

While the scheme outlined above imposes realistic diurnal- and synoptic-scale variations on monthly-mean GPP and RE, it still allows for abrupt changes from one month to the next. For CT2022, we add a further processing step designed to remove such unrealistic step changes. We fit smooth curves to the monthly GPP and RE using the piecewise integral quadratic splines (PIQS) of Rasmussen (1991). These PIQS fits are continuous in the first and second derivatives, and have the property of preserving monthly mean flux. We use a similar scheme to smooth over year-to-year step changes in fossil fuel emissions. The final smoothed GPP is

GPPF(t) = GPP(t) − GPPmean + GPPPIQS(t),
(8)
and the final smoothed ecosystem respiration is

RE,F(t) = RE(t) − RE,mean + RE,PIQS(t).
(9)
Together, these form the terrestrial NEE imposed as a first-guess flux in CT2022:

NEEF(t) = GPPF(t) + RE,F(t).
(10)
/webdata/ccgg/CT2022/summary/longterm_CT2022_flux1x1_glb_ltm.png
Figure 1: Map of optimized global biosphere fluxes. The pattern of net ecosystem exchange (NEE) of CO2 of the land biosphere averaged over the time period indicated, as estimated by CarbonTracker. This NEE represents land-to-atmosphere carbon exchange from photosynthesis and respiration in terrestrial ecosystems, and a contribution from fires. It does not include fossil fuel emissions. Negative fluxes (blue colors) represent CO2 uptake by the land biosphere, whereas positive fluxes (red colors) indicate regions in which the land biosphere is a net source of CO2 to the atmosphere. Units are g C m−2 yr−1.

2.3  GFED4.1s and GFED_CMS

CarbonTracker uses fluxes from CASA runs from two models associated with the GFED project as its first guess for terrestrial biosphere fluxes. We have found a significantly better match to observations when using this output compared to the fluxes from a neutral biosphere simulation. Both of the CASA simulations used in CT2022 (GFED 4.1s and GFED_CMS) are driven by AVHRR NDVI. This satellite driver tends to produce a larger-amplitude annual cycle of NEE compared to the alternative driver, MODIS fPAR. As one of the robust results of atmospheric inversions is a deeper annual cycle of terrestrial NEE, inversions using NDVI-driven first-guess fluxes perform slightly better than those with a MODIS fPAR driver.
The record of atmospheric CO2 calls for a deeper terrestrial biosphere sink than that generally simulated by terrestrial biosphere models like CASA. Inverse models manifesting such a sink generally simulate a larger annual cycle of terrestrial biosphere fluxes, and in particular a deeper boreal summer uptake of carbon dioxide, in the posterior optimized fluxes compared to the prior models (See Fig. 2). We call upon the atmospheric CO2 observations to make this change, and in order to handle these prior model differences the ensemble Kalman filter's prior covariance model has to be appropriately tuned. In short, this prior uncertainty needs to comfortably span differences among the terrestrial biosphere priors, the fossil fuel emissions estimates, and adjustments to fluxes required to bring model predictions into agreement with observations. CT2022 prior covariances have been adjusted compared to prior releases, and details on this adjustment can be found in Section 8.
/webdata/ccgg/CT2022/summary/land_global_totals.png
Figure 2: Time series of global-total terrestrial biosphere flux between the two priors and the CT2022 posterior. Global CO2 uptake by the land biosphere, expressed in Pg C yr−1, excluding emissions by wildfire. Positive flux represents emission of CO2 to the atmosphere, and the negative fluxes indicate times when the land biosphere is a sink of CO2. Optimization against atmospheric CO2 data requires a larger land sink than in either prior, which effectively requires a deeper annual cycle. This is shown by the CT2022 posterior (black).
/webdata/ccgg/CT2022/summary/dfb-glb.png
Figure 3: Differences in long-term mean terrestrial biosphere fluxes between the two priors. Red indicates areas where the GFED4.1s prior has less terrestrial uptake (or more outgassing to the atmosphere) than the GFED_CMS prior, and blue represents the opposite. Units are g C m−2 yr−1.
CarbonTracker CT2022 is a full reanalysis of the 2000-20w0 period using new fossil fuel emissions, CASA-GFED v4.1s and GFED_CMS fire emissions, and first-guess biosphere model fluxes derived from CASA-GFED v4.1s for the first of our inversions, and from CASA GFED_CMS for the second inversion.
Due to the inclusion of fires, inter-annual variability in weather and NDVI, the fluxes for North America start with a small net flux even before optimizing the fluxes. This first-guess flux ranges from neutral exchange to about 0.5 Pg C yr−1 of uptake.

3  Fire module

Vegetation fires are an important part of the carbon cycle and have been so for many millennia. Even before human civilization began to use fires to clear land for agricultural purposes, most ecosystems were subject to natural wildfires that would rejuvenate old forests and bring important minerals to the soils. When fires consume part of the landscape in either controlled or natural burning, carbon dioxide (among many other gases and aerosols) is released in large quantities. Each year, vegetation fires emit around 2 Pg C as CO2 into the atmosphere, mostly in the tropics. Currently, a large fraction of wildfire is started by humans. This is mostly intentional to clear land for agriculture, or to re-fertilize soils before a new growing season. This important component of the carbon cycle is monitored mostly from space, while sophisticated `biomass burning' models are used to estimate the amount of CO2 emitted by each fire. Such estimates are then used in CarbonTracker to prescribe the emissions. These emissions are not modified in the optimization (inverse modeling) process.
In CT2022 we use two fire emissions datasets, each with at least daily temporal resolution. The GFED4.1s emissions are modeled at 3-hourly intervals, and GFED_CMS emissions are available at daily resolution.

3.1  Global Fire Emissions Database (GFED)

CT2022 uses GFED4.1s (Giglio et al., 2013), (van der Werf et al., 2017) as one of the fire modules to estimate biomass burning. GFED4.1s is a variant of the CASA biogeochemical model as described in the terrestrial biosphere model documentation to estimate the carbon fuel in various biomass pools. The dataset consists of 1° × 1° gridded monthly burned area, fuel loads, combustion completeness, and fire emissions (Carbon, CO2, CO, CH4, NMHC, H2, NOx, N2O, PM2.5, Total Particulate Matter, Total Carbon, Organic Carbon, Black Carbon) for the time period spanning January 1997 - December 2021, of which we currently only use CO2.
The GFED burned area is based on MODIS satellite observations of fire counts. These, together with detailed vegetation cover information and a set of vegetation specific scaling factors, allow predictions of burned area over the time span that active fire counts from MODIS are available. The relationship between fire counts and burned area is derived, for the specific vegetation types, from a `calibration' subset of 500m resolution burned area from MODIS in the period 2001-2004.
Once burned area has been estimated globally, emissions of trace gases are calculated using the CASA biosphere model. The seasonally changing vegetation and soil biomass stocks in the CASA model are combusted based on the burned area estimate, and converted to atmospheric trace gases using estimates of fuel loads, combustion completeness, and burning efficiency.
For CT2022, we also apply temporal scaling factors updated from Mu et al. (2011) to downscale the GFED4.1s CO2 emissions from monthly averages to emissions with 3-hourly resolution.

3.2  GFED_CMS: Fluxes from the NASA Carbon Monitoring System

The NASA GFED_CMS team uses a variant of the GFED4 system to produce alternative fire emissions. This model uses GIMSS NDVI, the GFEDv3 fire model and GFEDv4 burned area. Fire emissions are available on a daily basis from 2003-2017. For 2000-2002, and for 2018-2020, we apply the climatology of GFED_CMS fire emissions, computed from its 2003-2017 mean.
Note that the GFED_CMS team produces temporally-downscaled GPP, heterotrophic respiration, and fires with 3-hourly resolution. This is done using MERRA meteorology using a scheme similar to Olsen and Randerson (2004). We do not use this downscaled product, in part because the MERRA meteorology is different from the ECMWF meteorology, and in part because the spatial resolution of the MERRA meteorology is different from our 1° × 1° flux grid. This means that we are limited to daily resolution of GFED_CMS fire emissions: unlike the GFED4.1s fire emissions, these have no diurnal cycle.

4  Fossil fuel module

Human beings first influenced the carbon cycle through land-use change. Early humans used fire to control animals and later cleared forests for agriculture. Over the last two centuries, following the industrial and technical revolutions and continuing global population increase, fossil fuel combustion has become the largest anthropogenic source of CO2. Coal, oil and natural gas combustion are the most common energy sources in both developed and developing countries. Global cement production is also significant, contributing about 5% of total fossil CO2 emissions. Important sectors of the economy-power generation, transportation, residential & commercial building heating, and industrial processes-rely on fossil fuels. The continued growth of fossil fuel combustion has led to a steady increase of global CO2 emissions to the atmosphere (Fig. 4). According to Boden et al. (2017), global emissions of CO2 from fossil fuel burning, cement manufacturing, and flaring reached 5 billion metric tons of carbon per year (Pg C yr−1) in the decade of the 1970s. Updated emissions products indicate that global total emissions exceeded 10 Pg C yr−1 for the first time in 2018. One petagram of carbon, Pg C, is equal to 1015 grams of carbon, or one billion metric tons of carbon. To convert to mass of CO2 emitted, one would multiply by the factor [44/12], representing the molecular weight of CO2 compared to the atomic weight of carbon.
/webdata/ccgg/CT2022/summary/CT2022_FF_tser.png
Figure 4: Time series of annual global fossil fuel emissions, in units of Pg C yr−1 (billion metric tons of C per year). Values from 1751 to 2014 are from Boden et al. (2017), and later values are extrapolated using consumption growth rate data of British Petroleum (2019). Inset figure shows the CT2022 period of analysis, 2000-2020.
U.S. input of CO2 to the atmosphere from fossil fuel burning in 2020 was 1.3 Pg C, representing 13% of the global total. North American emissions remained nearly constant from 2000-2018, and decreased slightly during the 2020 COVID pandemic year. On the other hand, emissions from developing economies such as the People's Republic of China have been increasing. Emissions from China in 2020 were 2.8 Pg C yr−1, representing 30% of the global total.
In almost all global and regional carbon flux estimation systems, including CarbonTracker, fossil fuel CO2 emissions are not optimized. Instead, these emissions are imposed and are not subject to revision by the inverse modeling framework. Global mass balance requires that any errors in fossil fuel emissions be compensated by opposing errors in land and ocean CO2 exchange. Thus it is vital that fossil fuel CO2 emissions are prescribed accurately, so that flux estimates for the land biosphere and oceans are robust. The fossil fuel emissions source data we use are available on an annually-integrated global and national basis. This aggregate information needs to be gridded before being incorporated into CarbonTracker. The major uncertainty in this process is distributing the national-annual emissions spatially across a nation and temporally into hourly contributions. In CT2022, two different fossil fuel CO2 emissions datasets were used to help assess the uncertainty in this mapping process. These two emissions products are called the "Miller" and "ODIAC" emissions datasets. These two datasets have very similar global and national emissions for each year, but differ in how those emissions are distributed spatially and temporally.
Whereas early CarbonTracker releases used monthly-constant fossil fuel emissions, starting with CT2015 we introduced the use of temporal scaling factors to simulate day-of-week and diurnal variability for those emissions. These "Temporal Improvements for Modeling Emissions by Scaling" (TIMES) scaling factors, introduced by Nassar et al. (2013), are again applied to both the Miller and ODIAC emissions modules for CT2022. The scaling factors consist of seven day-of-week global scaling factor maps, and 24 hourly global scaling factor maps to represent the diurnal cycle. For use in TM5, the hourly scaling factors were aggregated to three-hourly factors to accommodate the time step of the model.
/webdata/ccgg/CT2022/summary/longterm_CT2022_flux1x1_ff_glb_ltm.png
Figure 5: Spatial distribution of fossil fuel emissions. This is a spatial average of the Miller and ODIAC emissions inventories.

4.1  The "Miller" emissions dataset

  • Global and National Totals The Miller fossil fuel emission inventory is derived from independent global total and spatially-resolved inventories. Annual global total fossil fuel CO2 emissions are based on the Appalachian Engergy Center’s “CDIAC at AppState” project (https://energy.appstate.edu/research/work-areas/cdiac-appstate), which is an effort to update the original annual global and country fossil fuel-CO2 emissions estimates from the DOE’s Carbon Dioxide Information and Analysis Center (CDIAC) (Boden et al., 2017). The CDIAC at AppState emissions estimates used in CT2022 extend through 2017.
    In order to estimate these fluxes through April 2022 (the months in 2022 needed for our 12-week assimilation window), we extrapolate the CDIAC at AppState estimates using the fractional increases from other fossil fuel emissions products: 1) The BP (formerly British Petroleum) Statistical Review of World Energy (BP, 2021), and 2) The near-real-time CarbonMonitor product https://carbonmonitor.org/, accessed Dec. 6, 2022. To calculate emissions for 2018 and 2019, annual fractional emissions increases (by country and fuel type: coal, oil, and gas) are derived from BP data and applied to CDIAC at AppState data for 2017. For 2020 through mid-2022 emissions, fractional increases are applied to 2019 emissions estimates on a per-country and monthly basis. For example, to calculate July 2000 emissions for France, the ratio of July 2020 to July 2019 emissions for France are used to scale July 2019 emissions. No fuel-type data is available from CarbonMonitor, so the same year-on-year fractional changes are applied to all fuel types.
  • Spatial Distribution Miller fossil-fuel CO2 fluxes are spatially distributed in two steps: First, the coarse-scale country totals through 2017 from CDIAC at AppState are mapped onto a 1° × 1° grid according to the spatial patterns from the EDGAR v5.0 inventories (Commission et al., 2019). The spatial pattern varies by year up until the end of the EDGAR v5.0 product in 2018. After this, the trends estimated in each pixel are linearly extrapolated. Note that while EDGAR provides annual emissions estimates at 1° × 1° resolution, their totals do not agree with those from CDIAC at AppState. Thus, only the spatial patterns in EDGAR are used. The CDIAC country-by-country totals sum to about 95% of the global total emissions; the remaining 5% is mapped to global shipping routes according to EDGAR, which we treat as a proxy for bunker fuel emissions.
  • Temporal Distribution For North America between 30 and 60°N, the Miller system imposes a seasonal cycle derived from the first and second harmonics (Thoning et al., 1989) of the Blasing et al. (2004) analysis for the United States. The Blasing analysis has 10% higher emissions in winter than in summer. This scheme defines a fixed fraction of emissions for each month, so while the shape of the annual cycle is invariant, the amplitude of that cycle scales with the annual total emissions. For Eurasia, a set of seasonal emissions factors from EDGAR distributed by emissions sector is used to define fossil fuel seasonality. As in North America, this seasonality is imposed only from 30-60°N. The Eurasian seasonal amplitude is about 25%, significantly larger than that in North America, owing to the absence of a secondary summertime maximum due to air conditioning. See Figure 6 for the resulting time series of fossil fuel emissions. In order to avoid discontinuities in the fossil fuel emissions between consecutive years, a spline curve that conserves annual totals (Rasmussen, 1991) is fit to seasonal emissions in each 1° × 1° grid cell.

4.2  The "ODIAC" emissions dataset

  • Global Totals We use the ODIAC2020 fossil fuel emission inventory (Oda et al., 2018), (Oda and Maksyutov, 2011), (Oda and Maksyutov, 2015) as one of our fossil fuel emissions estimates. This version was prepared for use in the OCO-2 v10 MIP (Basu and Nassar, 2021), (Byrne et al., 2022) and includes a 2020 anomaly derived from the Carbon Monitor emissions project (https://carbonmonitor.org/). ODIAC is also derived from independent global and country emission estimates from CDIAC, but national emission estimates used were taken from (Gilfillan and Marland, 2021). Differences between the Gilfillan and Marland (2021) global total and country-by-country totals were ascribed to the entire emissions field. Annual country total fossil fuel CO2 emissions for 2018-2019 were extrapolated using the BP Statistical Review of World Energy (British Petroleum, 2019), and those for 2020 used the Carbon Monitor results as described in Byrne et al. (2022).
  • Spatial Distribution ODIAC emissions are spatially distributed using many available "proxy data" that explain spatial extent of emissions according to emission types (emissions over land, gas flaring, aviation and marine bunker). Emissions over land were distributed in two steps: First, emissions attributable to power plants were mapped using geographical locations (latitude and longitude) provided by the global power plant dataset CARbon Monitoring and Action, CARMA. Next, the remaining land emissions (i.e. land total minus power plant emissions) were distributed using nightlight imagery collected by U.S. Air Force Defense Meteorological Satellite Project (DMSP) satellites. Emissions from gas flaring were also mapped using nightlight imagery. Emissions from aviation were mapped using flight tracks adopted from UK AERO2k air emission inventory. It should be noted that currently, air traffic emissions are emitted at ground level within CarbonTracker. Emissions from marine bunker fuels are placed entirely in the ocean basins along shipping routes according to patterns from the EDGAR database.
  • Temporal Distribution The CDIAC estimates used for mapping emissions in ODIAC only describe how much CO2 was emitted in a given year. To present seasonal changes in emissions, we used the CDIAC 1° × 1° monthly fossil fuel emission inventory (Andres et al., 2011). The CDIAC monthly data utilizes the top 20 emitting countries' fuel (coal, oil and gas) consumption statistics available to estimate seasonal change in emissions. Monthly emission numbers at each pixel were divided by annual total and then a fraction to annual total was obtained. Monthly emissions in the ODIAC inventory were derived by multiplying this fraction by the emission in each grid cell.
/webdata/ccgg/CT2022/summary/ff_global_totals.png
Figure 6: Time series of global fossil fuel emissions showing annual cycles. The Miller (green) and ODIAC (tan) estimates are each used by half of the sixteen inversions in the CT2022 suite, so the CT2022 (black) inventory is effectively an average of Miller and ODIAC. Note that fossil fuel emissions are not optimized in CarbonTracker.
/webdata/ccgg/CT2022/summary/dff-nam.png
/webdata/ccgg/CT2022/summary/dff-asi.png
Figure 7: Spatial differences in long-term mean fossil fuel emissions between the two priors. Note that both the Miller and ODIAC emissions inventories use the same country totals, but have different models for spatial distribution of that flux within countries.

4.3  Uncertainties

Marland (2008) attached an uncertainty of about 5% (95% confidence interval; approximately 2-σ) to the global total fossil fuel source. Estimates by Andres et al. (2014) put a larger uncertainty of 8.4% (2-σ) on the CDIAC global total. Uncertainties for individual regions of the world, and for sub-annual time periods are likely to be larger. Additional uncertainties are introduced when the emissions are distributed in space and time. In the Miller dataset, the overall Eurasian seasonality is based on scaling factors derived only from Western Europe and thus highly uncertain, but most likely a better representation than assuming no emission seasonality at all. Similarly, the use of the CDIAC monthly emission dataset for modeling seasonality introduces additional uncertainty in ODIAC. The additional uncertainty for the global total in the monthly CDIAC emission, which is solely due to the method for estimating seasonality, is reported as 6.4% (Andres et al., 2011). As mentioned earlier, fossil fuel emissions are not optimized in the current CarbonTracker system, similar to nearly all carbon data analysis systems. Spatial and temporal atmospheric CO2 gradients arise from terrestrial biosphere and fossil-fuel sources. These gradients, which are interpreted by CarbonTracker, are difficult to attribute to one or the other cause. This is because atmospheric sampling sites have historically been established in locations remote from biospheric and anthropogenic sources, especially in the temperate Northern Hemisphere. Given that surface CO2 flux due to biospheric activity and oceanic exchange is much more uncertain compared to fossil fuel emissions, CarbonTracker, like most current carbon dioxide data assimilation systems, does not attempt to optimize fossil fuel emissions. That is, the contribution of CO2 from fossil fuel burning to observed CO2 mole fractions is considered known. As detailed above, however, in CarbonTracker an effort is made to account for some aspects of fossil fuel uncertainty by using two different fossil fuel estimates. From a technical point of view, extra land biosphere prior flux uncertainty is included in the system to represent the random errors in fossil fuel emissions. Eventually, fossil fuel emissions could be optimized within CarbonTracker, especially with the addition of 14CO2 observations as constraints (Basu et al., 2016), (Basu et al., 2020).

5  Oceans module

The oceans play an important role in the Earth's carbon cycle. They are the largest long-term sink for carbon and have an enormous capacity to store and redistribute CO2 within the Earth system. Oceanographers estimate that about 48% of the CO2 from fossil fuel burning has been absorbed by the ocean (Sabine et al., 2004). The dissolution of CO2 in seawater shifts the balance of the ocean carbonate equilibrium towards a more acidic state with a lower pH. This effect is already measurable (Caldeira and Wickett, 2003), and is expected to become an acute challenge to shell-forming organisms over the coming decades and centuries. Although the oceans as a whole have been a relatively steady net carbon sink, CO2 can also be released from oceans depending on local temperatures, biological activity, wind speeds, and ocean circulation. These processes are all considered in CarbonTracker, since they can have significant effects on the ocean sink. Improved estimates of the air-sea exchange of carbon in turn help us to understand variability of both the atmospheric burden of CO2 and terrestrial carbon exchange.
/webdata/ccgg/CT2022/summary/longterm_CT2022_flux1x1_ocn_ltm.png
Figure 8: Posterior long-term mean ocean fluxes from CarbonTracker. The pattern of air-sea exchange of CO2 averaged over the time period indicated, as estimated by CarbonTracker. Negative fluxes (blue colors) represent CO2 uptake by the ocean, whereas positive fluxes (red colors) indicate regions in which the ocean is a net source of CO2 to the atmosphere. Units are g C m−2 yr−1.
The initial release of CarbonTracker (CT2007) used climatological estimates of CO2 partial pressure in surface waters (pCO2) from Takahashi et al. (2002) to compute a first-guess air-sea flux. This air-sea pCO2 disequilibrium was modulated by a surface barometric pressure correction before being multiplied by a gas-transfer coefficient to yield a flux. Starting with CT2007B and continuing through the CT2011_oi release, the air-sea pCO2 disequilibrium was imposed from analysis of ocean inversions (Jacobson et al., 2007),"OIF" results, with short-term flux variability derived from the atmospheric model wind speeds via the gas transfer coefficient. The barometric pressure correction was removed so that climatological high- and low-pressure cells did not bias the long-term means of the first guess fluxes.
In CT2022, two models are used to provide prior estimates of air-sea CO2 flux. The OIF scheme provides one of these flux priors, and the other is an updated version of the Takahashi et al. (2009) pCO2 climatology.

5.1  Air-sea gas exchange

Oceanic uptake of CO2 in CarbonTracker is computed using air-sea differences in partial pressure of CO2 inferred either from ocean inversions (called "OIF" henceforth), or from a compilation of direct measurements of seawater pCO2 (called "pCO2-clim" henceforth). These air-sea partial pressure differences are combined with a gas transfer velocity computed from wind speeds in the atmospheric transport model to compute fluxes of carbon dioxide across the sea surface.
In either method, the first-guess fluxes have no interannual variability (IAV) other than a smooth trend. IAV in oceanic CO2 flux is due to anomalies in surface pCO2, such as those that occur in the tropical eastern Pacific during an El Niño, and to associated variability in winds, ocean circulation, and sea-surface properties. In CarbonTracker, only the surface winds (and hence gas transfer), manifest these interannual anomalies; the remaining IAV of flux must be inferred from atmospheric CO2 signals.
In the following sections we describe the two ocean flux prior models. We then describe the air-sea gas transfer velocity parameterization and discuss details of the inversion methodology specific to oceanic exchange of CO2.

5.2  OIF: the Ocean Inversion Fluxes prior

For the OIF prior, long-term mean air-sea fluxes and the uncertainties associated with them are derived from the ocean interior inversions reported in Jacobson et al. (2007). These ocean inversion flux estimates are composed of separate preindustrial (natural) and anthropogenic flux inversions based on the methods described in Gloor et al. (2003) and biogeochemical interpretations of Gruber et al. (1996). The uptake of anthropogenic CO2 by the ocean is assumed to increase in proportion to atmospheric CO2 levels, consistent with estimates from ocean carbon models.
OIF contemporary pCO2 fields were computed by summing the preindustrial and anthropogenic flux components from inversions using five different configurations of the Princeton/GFDL MOM3 ocean general circulation model (Pacanowski and Gnanadesikan, 1998), then dividing by a gas transfer velocity computed from the European Centre for Medium-Range Weather Forecasts (ECMWF) ERA40 reanalysis. There are two small differences in first-guess fluxes in this computation from those reported in Jacobson et al. (2007). First, the five OIF estimates all used Takahashi et al. (2002) pCO2 estimates to provide high-resolution patterning of flux within inversion regions (the alternative "forward" model patterns were not used). To good approximation, this choice only affects the spatial and temporal distribution of flux within each of the 30 ocean inversion regions, not the magnitude of the estimated flux. Second, wind speed differences between the ERA40 product used in the offline analysis and the ECMWF operational model used in the online CarbonTracker analysis result in small deviations from the OIF estimates.
Other than the smooth trend in anthropogenic flux assumed by the OIF results, interannual variability (IAV) in the first guess ocean flux comes entirely from wind speed effects on the gas transfer velocity. This is because the ocean inversions retrieve only a long-term mean and smooth trend.

5.3  pCO2-Clim: Takahashi climatology prior

The pCO2-Clim prior is derived from the Takahashi et al. (2009) climatology of seawater pCO2. This climatology was created from about 3 million direct observations of seawater pCO2 around the world between 1970 and 2007. With the exception of measurements in the Bering Sea, these observations were all linearly extrapolated to the corresponding month of the year 2000 by assuming a constant trend of 1.5  μatm yr−1. This set of global monthly measurements corrected to the reference year 2000 was then interpolated onto a regular grid using a modeled surface current field.
The Takahashi et al. (2009) product goes beyond providing this estimate of surface water pCO2. They also compute climatological air-sea exchange of CO2 by using the GLOBALVIEW-CO2 atmospheric carbon dioxide product to compute air-sea ∆pCO2, sea surface properties inferred from ocean climatologies, and winds from atmospheric reanalysis to estimate gas-transfer velocity. Unlike many other atmospheric analyses, we have chosen not to use the climatological fluxes as our prior, nor to use the climatological ∆pCO2. Instead, we take only the seawater pCO2 distribution from the Takahashi et al. (2009) climatology-our atmospheric model provides both pCO2 in the air at the sea surface and the winds needed to estimate gas transfer. Seawater pCO2 is extrapolated from 2000 to the actual year of the CarbonTracker simulation using a presumed increase of 1.5  μatm yr−1 at every point in the global ocean. This is the same trend used in Takahashi et al. (2009) to normalize observations from many years to the reference year of the analysis (2000).

5.4  Gas-transfer velocity and ocean surface properties

Both priors use CO2 solubilities and Schmidt numbers computed from World Ocean Atlas 2009 (WOA09) climatological fields of sea surface temperature and sea surface salinity fields (Levitus et al., 2010). Gas transfer velocity in CarbonTracker is parameterized as a quadratic function of wind speed following Wanninkhof (1992), using the formulation for instantaneous winds. Gas exchange is computed every 3 hours using wind speeds from the ECMWF operational model as represented by the atmospheric transport model.
Air-sea transfer is inhibited by the presence of sea ice, and for this work fluxes are scaled by the daily sea ice fraction in each gridbox provided by the ECMWF forecast data.
/webdata/ccgg/CT2022/summary/ocean_global_totals.png
Figure 9: Comparison of air-sea flux priors and the CT2022 posterior. Global CO2 uptake by the ocean, expressed in Pg C yr−1. Positive flux represents a gain of CO2 to the atmosphere, and the negative numbers here indicate that the ocean is a sink of CO2. While both priors manifest similar trends of increasing oceanic uptake of CO2, the OIF prior (in green) has more oceanic uptake and a greater annual cycle than the pCO2-clim prior (in tan). The CT2022 across-model posterior estimate is shown in black for comparison.
/webdata/ccgg/CT2022/summary/dfo-glb.png
Figure 10: Differences in long-term mean ocean fluxes between the two priors. Red indicates areas where the pCO2-clim prior has less oceanic uptake (or more outgassing to the atmosphere) than the OIF prior, and blue represents the opposite. Units are gC m-2 yr-1.

5.5  Specifics of the inversion methodology related to air-sea CO2 fluxes

The first-guess fluxes described here are subject to scaling during the CarbonTracker optimization process, in which atmospheric CO2 mole fraction observations are combined with transport simulated by the atmospheric model to infer flux signals. Prior air-sea fluxes are adjusted within each of the 30 ocean inversion regions. In this process, signals of terrestrial flux in atmospheric CO2 distribution can be erroneously interpreted as being caused by oceanic fluxes. This flux "aliasing" or "leakage" is evident in some regions as a change in the shape of the seasonal cycle of air-sea flux.
Prior uncertainties for the OIF and pCO2-clim models are specified as uncertainties on scaling factors multiplying net CO2 flux in each of the 30 ocean inversion regions. The pCO2-clim prior has independent regional uncertainties (a diagonal prior covariance matrix), with the uncertainty standard deviation on each region set to 40%. The OIF prior uncertainty has a fully-covariate covariance matrix with off-diagonal elements representing the results of the ocean inversion of Jacobson et al. (2007). The preindustrial flux uncertainty is time-independent, but the anthropogenic flux uncertainty grows in time as anthropogenic flux uptake increases. The latter is scaled to the simulation date, then added to the former. Total uncertainties are consistent with the Jacobson et al. (2007) results.

6  Atmospheric transport

The link between observations of CO2 in the atmosphere and the exchange of CO2 at the Earth's surface is transport in the atmosphere: storm systems, cloud complexes, and weather of all sorts cause winds that transport CO2 around the world. As a result, local surface CO2 exchange events like fires, forest growth, and ocean upwelling can have impacts at remote locations. To simulate the winds and the weather, CarbonTracker uses sophisticated numerical models that are driven by the daily weather forecasts from the specialized meteorological centers of the world. Since CO2 does not decay or react in the lower atmosphere, the influence of emissions and uptake in locations such as North America and Europe are ultimately seen in our measurements even at the South Pole. Getting the transport of CO2 just right is an enormous challenge, and costs us almost all of the computer resources for CarbonTracker. To represent the atmospheric transport, we use the Transport Model 5 (TM5). This is a community-supported model whose development is shared among many scientific groups with different areas of expertise. TM5 is used for many applications other than CarbonTracker, including forecasting air-quality, studying the dispersion of aerosols in the tropics, tracking biomass burning plumes, and predicting pollution levels that future generations might have to deal with.

6.1  TM5 offline tracer transport model

TM5 is an offline global chemical transport model with two-way nested grids. In this global model, regions for which high-resolution simulations are desired can be nested in the coarser global grid. The advantage to this approach is that transport simulations can be performed with a regional focus without the need for boundary conditions. Further, this approach allows measurements outside the "zoom" domain to constrain regional fluxes in the data assimilation, and ensures that regional estimates are consistent with global constraints. TM5 is based on a predecessor model TM3, with improvements in the advection scheme, vertical diffusion parameterization, and meteorological preprocessing of the wind fields (Krol et al., 2005).
The model is developed and maintained jointly by the Institute for Marine and Atmospheric Research Utrecht (IMAU, The Netherlands), the Joint Research Centre (JRC, Italy), the Royal Netherlands Meteorological Institute (KNMI), the Netherlands Institute for Space Research (SRON), and the NOAA Global Monitoring Laboratory (GML).
In CarbonTracker, TM5 separately simulates advection, deep and shallow convection, and vertical diffusion in both the planetary boundary layer and free troposphere. The carbon dioxide concentrations predicted by CarbonTracker do not feed back onto these predictions of winds.
Prior to use in TM5, ECMWF meteorological data are preprocessed into coarser grids, with attention to retrieving a flow that conserves tracer mass. Like most numerical weather prediction models, advection in the parent ECMWF model is not strictly mass-conserving, so this step is crucial. In CarbonTracker, TM5 is currently run at a global 3° longitude × 2° latitude resolution with a nested regional grid over North America at 1° × 1° resolution (Figure 11). TM5 uses a dynamically-variable time step with a maximum length of 90 minutes. This overall timestep is dynamically reduced to maintain numerical stability, generally during times of high wind speeds. The timestep is divided in half and individual advection, diffusion, convection, and chemistry operators are applied symmetrically in each half step. Furthermore, transport operators in nested grids are modeled at shorter timesteps, so processes at the finest scales are conducted at an effective timestep of one-quarter the overall timestep. See Krol et al. (2005) for details.
images/CT_grids.png
Figure 11: Nested grids used in CarbonTracker over North America. TM5 is a global model, but it employs nested grids to provide higher resolution over regions of interest. This figure shows the 1° × 1° nested regional grid over North America and a portion of the global 3° longitude × 2° latitude grid.
The winds which drive TM5 come from the ERA5 reanalysis implemented in the European Centre for Medium-Range Weather Forecasts (ECMWF) modeling system. The ERA5 reanalysis uses CY41R2 version of the ECMWF Integrated Forecast System (IFS) model. That model uses a 12-minute time step and a spectral T639 horizontal resolution, which corresponds to approximately 28 km spacing at the equator on a reduced Gaussian grid. This version of the IFS has 137 model layers in the vertical, of which TM5 uses a 34-layer subset. These levels are listed in Table 1.
Model Level Mean Height (m) Model Level Mean Height (m)
1 33 18 9400
2 109 19 10131
3 255 20 11011
4 477 21 11749
5 814 22 12492
6 1273 23 13393
7 1835 24 14304
8 2556 25 15226
9 3315 26 16322
10 4205 27 17446
11 5026 28 18459
12 5603 29 20380
13 6186 30 24376
14 6771 31 29834
15 7355 32 35623
16 8086 33 42602
17 8816 34 123210
Table 1: Mean mid-level heights in meters above ground from the ERA5 reanalysis using the TM5 34 level subset.

6.2  Convective flux fix

Until recently, TM5 was known to have difficulties representing the global surface distribution of sulfur hexafluoride (SF6, see Figure 12 and Peters et al. (2004)). SF6 is a nearly inert tracer in the atmosphere, with very small surface and atmospheric sinks and an atmospheric lifetime of about 1,000 years. Consequently, its global budget is very well known from observations alone. It is thought to be released mainly via leakage from electrical transformers. Since the electrical distribution system is closely tied to fossil fuel consumption, SF6 is often considered an analog for fossil fuel CO2 in the atmosphere. It is useful for understanding the rate at which Northern Hemisphere land surfaces are ventilated to the free troposphere, and the rate of interhemispheric exchange in models (Patra et al., 2011).
As a result of more than a decade's worth of work on understanding the apparently sluggish mixing in TM5 as revealed by SF6 simulations, a fault in one of the vertical mixing parameterizations of the model was discovered. When it was originally created, TM5 implemented the same planetary boundary layer (PBL) mixing and convection schemes as the parent ECMWF model. Recent comparisons between TM5, the ECMWF parent model, and radiosonde profile data show that the PBL scheme in TM5 performs similarly to that of the parent ECMWF model. The convective scheme, however, does not produce similar results in TM5 as compared to the ECMWF model.
/webdata/ccgg/CT2022/summary/sf6_latscat_1.png
Figure 12: Long-term mean model residuals of SF6 concentrations as a function of latitude. Residuals are defined as model-minus-observation, so a positive residual indicates the model has too much SF6. Three different transport model simulations are shown. The ECMWF forecast (blue) and ERA-interim (red) transport simulations do not include the recent "convective flux fix". The ERA-interim with this convective flux fix is shown in green. Units are  pmol mol−1, or parts per trillion. CT2022 uses the ERA-interim transport with the convective flux fix.
In a previous configuration of TM5, the convective entrainment and detrainment mass fluxes of the parent ECMWF model were re-diagnosed within TM5 using other meteorological information. The ECMWF model is used to produce both operational forecasts and the ERA-interim reanalysis, but the convective fluxes are stored for the ERA-interim product only. Thus, using ERA-interim meteorology, a direct comparison is possible. This comparison revealed that the TM5 internal rediagnosis of convective fluxes was faulty. TM5 was subsequently modified to use parent model ERA-interim convective fluxes directly. Using the parent model convective fluxes result in a significantly better SF6 simulations. Simulations with these parent-model convective fluxes are said to use the "convective flux fix". Simulations with the convective flux fix show significantly improved agreement with SF6 observations (see Figure 12).
Since the parent-model convective fluxes are only available for the ERA-interim product, CT2022 uses only ERA-interim transport with the convective flux fix. Previous releases of CarbonTracker also used the ECMWF operational model transport, for which parent-model convective fluxes are not available. We believe that TM5 simulations without the parent-model convective fluxes are faulty and should not be included in our product. When the convective flux fix was instituted in CT2013B, it resulted in the largest realignment of surface CO2 fluxes in the history of the CarbonTracker program (Schuh et al., 2019). This is a prominent example of the sensitive reliance of atmospheric inversions on accurate atmospheric transport.

7  Observations

The observations of atmospheric CO2 mole fraction made by NOAA GML and partner laboratories are at the heart of CarbonTracker. They inform us on changes in the carbon cycle, whether those changes are regular (such as the annual cycle of growth and decay of leaves and other plant matter), or irregular (such as the release of tons of carbon by a wildfire). The results in CarbonTracker depend directly on the quality, location, and frequency of available observations. The level of detail at which we can retrieve information on the carbon cycle increases strongly with the density of the CO2 observing network.

7.1  The CarbonTracker observational network

Observations simulated by CT2022 are supplied by the GLOBALVIEW+ data product version 7.0 , available at the NOAA GML ObsPack web site. This study uses measurements of air samples collected at 559 sites around the world by 66 laboratories:
  • Penn State University (PSU)
  • NOAA Global Monitoring Laboratory (NOAA)
  • Instituto de Pesquisas Energeticas e Nucleares (IPEN)
  • Environment and Climate Change Canada (ECCC)
  • AVOCET Group @ NASA LaRC (NASA-LaRC)
  • National Institute for Environmental Studies (NIES)
  • CSIRO Oceans and Atmosphere, Climate Science Centre - GASLAB (CSIRO)
  • Chinese Academy of Meteorological Sciences (CMA)
  • Scripps Institution of Oceanography (SIO)
  • Scripps Institution of Oceanography CO2 Program (SIO_CO2)
  • Max Planck Institute for Biogeochemistry (MPI-BGC)
  • Laboratoire des Sciences du Climat et de l'Environnement - UMR8212 CEA-CNRS-UVSQ (LSCE)
  • Japan Meteorological Agency (JMA)
  • NOAA Chemical Sciences Division (NOAA-CSD)
  • National Institute of Water and Atmospheric Research (NIWA)
  • ICOS ATMOSPHERE THEMATIC CENTRE (ICOS-ATC)
  • Norwegian Institute for Air Research (NILU)
  • University of Bern, Physics Institute, Climate and Environmental Physics (KUP)
  • Atmospheric Chemistry Research Group School of Chemistry University of Bristol (UNIVBRIS)
  • Harvard University (HU)
  • Netherlands Organisation for Applied Scientific Research (TNO)
  • California Institute of Technology, Division of Geological and Planetary Science (CALTECH)
  • Institute of Atmospheric Sciences and Climate (CNR-ISAC) (CNR-ISAC)
  • Meteorological Research Institute (MRI)
  • South African Weather Service (SAWS)
  • University of Exeter, Centre for Environmental Data Analysis (CEDA)
  • Institut de Ciencia i Tecnologia Ambientals, Universitat Autonoma de Barcelona (ICTA-UAB)
  • Lawrence Berkeley National Laboratory and ARM Climate Research Facility (LBNL-ARM)
  • Hohenpeissenberg Meteorological Observatory (HPB)
  • Earth Networks, Inc. (EN)
  • NASA Goddard Space Flight Center (NASA-GSFC)
  • National Center For Atmospheric Research (NCAR)
  • University of Heidelberg, Institut fuer Umweltphysik (UHEI-IUP)
  • NOAA ESRL Halocarbons and Other Atmospheric Trace Species (NOAA-HATS)
  • Hong Kong Observatory (HKO)
  • Lund University - Centre for Environmental and Climate Research (LUND-CEC)
  • Hungarian Meteorological Service (HMS)
  • Karlsruhe Institute of Technology (IMK-ASF) (KIT/IMK-ASF)
  • Institute for Atmospheric and Environmental Sciences, University of Frankfurt (IAU)
  • Joint Research Centre (JRC)
  • Izana Atmospheric Research Center, Meteorological State Agency of Spain (AEMET)
  • High Altitude Research Stations Jungfraujoch and Gornergrat International Foundation (HFSJG)
  • Swiss Federal Laboratories for Materials Science and Technology (EMPA)
  • University of Science and Technology (AGH) (AGH)
  • University of Minnesota (UofMN)
  • Integrated Carbon Observation System - Flask and Calibration Laboratory (ICOS)
  • Czechglobe - Global Change Research Institute CAS (CAS)
  • University of Wisconsin (UofWI)
  • National Agency for New Technology, Energy, and Environment (ENEA)
  • University of Groningen (RUG), Centre for Isotope Research (CIO) (RUG)
  • Oregon State University (OSU)
  • Finnish Meteorological Institute (FMI)
  • Ricerca sul Sistema Energetico (RSE)
  • Commissariat à l'énergie atomique et aux énergies alternatives (CEA)
  • Savannah River National Laboratory (SRNL)
  • University of Helsinki (UHELS)
  • University of Virginia (UofVA)
  • Umweltbundesamt, Station Schauinsland (UBA-SCHAU)
  • Lawrence Berkeley National Laboratory (LBNL)
  • Forest Ecology and Management, SLU Umeå (SLU)
  • Center for Atmospheric and Oceanic Studies, Tohoku University (TU)
  • University of East Anglia (UEA)
  • Main Geophysical Observatory (MGO)
  • Utah Atmospheric Trace gas & Air Quality (U-ATAQ)
  • Umweltbundesamt, Zugspitze GAW Station (UBA/ZUG)
The CO2 measurement data assimilated in CT2022 are freely available for download from the GML ObsPack web portal or from partner websites. The bulk of assimilated measurements come from GLOBALVIEWplus v7.0 (2021) and from the GLOBALVIEWplus v8.0 product. Additional observations were gathered from specialized ObsPack products as detailed in Table 2.
Source Online availability Comment
GLOBALVIEW+ v7.0 ObsPack download Main source of observations
GLOBALVIEW+ v8.0 ObsPack download Observations since Jan 1, 2021
NIES shipboard observations Available from NIES No permission to redistribute
Table 2: Sources of CO2 observational data for CT2022
We also make available an ObsPack containing the simulated values of all measurement data considered by CT2022. This CT2022 ObsPack contains most, but not all, of the measured values. Measured values are only distributed directly when we have permission to do so.
Users are encouraged to review the usage requirements for these data products, and to contact the measurement laboratories directly for details about the observations.
/webdata/ccgg/CT2022/summary/network-nam.png
Figure 13: CarbonTracker observational network over North America. See the CarbonTracker interactive network map for more details.
With the advent in 2015 of GLOBALVIEW+, data are now presented to CarbonTracker with a higher temporal frequency than in past observational products. At sites with quasi-continuous monitoring, CT2022 assimilates hourly average CO2 concentrations. In the past, a single daily assimilation value was constructed at these sites, generally a four-hour average during well-mixed background conditions. At continental sites, this four-hour period was generally from local noon to 4pm; at many mountain sites background conditions are met at nighttime when upslope winds are uncommon. Using GLOBALVIEW+, CarbonTracker can now assimilate each hourly average during these background conditions independently. For many sites, all available hourly averages throughout the day are assimilated. Details vary by dataset, but can be checked at the interactive data plotting page.
Note that all of these observations are calibrated against the same world CO2 standard (WMO-X2007).
Starting with GLOBALVIEW+, we generally use the recommendations of data providers as to which observations are appropriate for assimilation. Such observations are identified by a variable in the ObsPack distribution, obs_flag. Only observations with obs_flag = 1 are identified for assimilation by data providers. We modify the designation of assimilation data for Environment and Climate Change Canada quasi-continuous sampling sites. For these data, obs_flag is set to 1 by the data provider for times when they represent the daily minimum CO2 concentration. This is generally later in the day than our standard scheme of local noon-4pm used to represent times of well-mixed PBLs. For these datasets, we have changed obs_flag to indicate assimilation only for the local noon - 4pm time period. These selected observations are further filtered based on the CCG curve fitting routine of Thoning et al. (1989). This filter fits a smooth curve to the selected observations, and measurements more than 3 standard deviations away from this curve are excluded from assimilation.
At mountain-top sites (e.g. MLO, NWR, and SPL), it is usually nighttime hours that are selected for assimilation, as these tend to be the most stable time period. Nighttime hours also avoid periods of upslope flows that contain local vegetative and/or anthropogenic influence.
Data from the Sutro tower (STR) and the Boulder (Erie, Colorado) tower (BAO) are strongly influenced by local urban emissions, which CarbonTracker is unable to resolve. At these two sites, pollution events have been identified using co-located measurements of carbon monoxide. In this study, measurements thought to be affected by pollution events have been excluded. This technique is under active refinement.
With CT2022, we have begun to assimilate CO2 measurements from NOAA light aircraft profiling time series, from intakes at multiple levels on NOAA tall towers, and from extensive shipboard and Siberian tower measurements collected by our partners at NIES. These datasets can be explored at the interactive data plotting page.
/webdata/ccgg/CT2022/summary/network-global.png
Figure 14: CarbonTracker global observational network. See the CarbonTracker interactive network map for more details.
We apply a further selection criterion during the assimilation to exclude non-marine boundary layer (MBL) observations that are very poorly forecasted in our framework. We use the so-called model-data mismatch in this process, which is the random error ascribed to each observation to account for measurement errors as well as modeling errors of that observation. We interpret an observed-minus-forecasted mole fraction that exceeds 3 times the prescribed model-data mismatch as an indicator that our modeling framework fails. This can happen for instance when an air sample is representative of local exchange not captured well by our 1° × 1° fluxes, when local meteorological conditions are not captured by our offline transport fields, but also when large-scale CO2 exchange is suddenly changed (e.g. fires, pests, droughts) to an extent that can not be accommodated by our flux modules. This last situation would imply an important change in the carbon cycle and has to be recognized by the researchers when analyzing the results. In accordance with the 3-sigma rejection criterion, about 0.2% of the observations are discarded through this mechanism in our assimilation.

7.2  Adaptive model-data mismatch

The statistical optimization method we use to constrain surface CO2 fluxes requires that each assimilation constraint is assigned a "model-data mismatch" (MDM) error value. This is meant to express the statistics of simulated-minus-observed CO2 observations we could expect if CarbonTracker were using perfect surface fluxes. Such deviations arise from many sources, including random noise in the measurement system, in situ variability that we do not expect to resolve in our model, and faults with the atmospheric transport model. Generally, transport and inverse model faults are the dominant terms in MDM values. The MDM is one of two major "tuning knobs" used to adjust the performance of our ensemble Kalman filter. The other is also an error quantity, meant to represent the expected error on our first-guess fluxes. Discussion of this prior covariance error can be found in section 8.2.
Prior to CT2015, CarbonTracker used a single MDM value for each assimilation dataset. The NOAA continuous observations at the 396m level of the WLEF tower in northern Wisconsin, for example, were assigned a MDM of 3.0 ppm, meaning that the residuals between model-forecasted measurements and the actual observed concentrations are expected to be unbiased (i.e., have a mean of zero) and have a standard deviation of 3 ppm. In practice, however, we have found that it is far easier to simulate wintertime observations than those during summer. This is mainly due to higher ambient variability of CO2 in the summer.
Starting with CT2016, we began to use an empirical scheme to assign MDM values, exploiting statistics of model performance from independently-configured preliminary inversions. The posterior residuals for each dataset are classified into relevant bins, and then statistics of model performance are analyzed within each of those bins. For every dataset, these bins include equally-spaced intervals of one-tenth of a year. For analyzers collecting data throughout the day, we also classify the measurements into 4-hour intervals of local time. For aircraft datasets, we further classify measurements into vertical levels of 1000m thickness (0-1000 m ASL, 1000-2000 m ASL, etc.). For each of these bins, bias and random error are combined to form total deviation from observed values as a root-mean square error (RMSE). The assigned MDM is set to a constant fraction of this total RMSE. This scaling is meant to force the assimilation scheme to extract as much information as possible from available observations. We use two different scaling factors to convert RMSE to MDM, depending on whether the preliminary inversions actually assimilated the measurements in the relevant bin, or merely simulated those measurements. For measurements assimilated by the preliminary inversions, the MDM is 0.95 * RMSE; for measurements not assimilated in the preliminary inversions, the MDM is 0.85 * RMSE.
The adaptive MDM scheme performs well in terms of average χ2, which in an optimally-tuned system should be close to 1.0 for each dataset (see Table 3). Notably, the seasonal variations of MDM successfully compensate for the higher ambient variability of CO2 at continental sites during the growing season. It is, however, an iterative process, requiring that we conduct a previous inversion. For various reasons, this previous inversion performed before CT2022 differs in significant aspects from the actual CT2022 inversions. These differences have led to MDM values which are slightly too large and thus average χ2 values which are generally smaller than the target of 1.0 (in some cases, as low as 0.2 or 0.3). The next iteration of CarbonTracker will be able to use the more recent CT2022 inversions to refine the adaptive MDM scheme.
Duplicate observations are identified as those within 50 minutes temporally, 10m vertically, and 0.05 degrees of latitude and longitude laterally (nominally, about 5km). The MDM for such observations is inflated by √n, where n is the number of duplicates.

7.3  Statistical performance of CT2022

Starting with CT2022, we reserved about 5% of available assimilation data for a cross-validation exercise. To the extent possible, withheld measurements were chosen to be independent from other observations. For surface flask observations, which are generally collected on a weekly time basis, all samples are considered independent from one another, so withheld data were selected randomly. Aircraft flasks collected during a profile are considered co-dependent, so entire profiles were randomly selected for withholding. Finally, for in situ analyzers with quasi-continuous sampling (towers, observatories), 24-hour periods were randomly chosen and that entire day's worth of data were withheld. Shipboard quasi-continuous datasets, which account for more than a quarter of assimilation data for CT2022 were inadvertently excluded from this cross-validation exercise. As a result, about 3% (115,756 measurements out of about 4.2 million) were withheld.
Each residual from the withheld measurements was normalized by its prescribed model-data mismatch (MDM) error, to form a set of χ values. The mean χ2, which for a perfect set of independent, normally-distributed variates should approach 1.0, was found to be 0.92.
Table 3 summarizes the datasets assimilated in CarbonTracker, and the performance of the assimilation scheme for each dataset. These diagnostics are useful for evaluating how well CarbonTracker does in simulating observed CO2.
Dataset Lab. Location Latitude Longitude Elev. Used Rej. Unsampled R χ2 Bias SE
(m) (ppm) (ppm) (ppm)
co2_abp_surface-flask_1_representative
NOAA Arembepe, Bahia, Brazil 12.77°S 38.17°W 1 90 1 0 0.1 - 4.5 0.41 -0.98 0.97
co2_abp_surface-flask_26_marine IPEN Arembepe, Bahia, Brazil 12.77°S 38.17°W 1 100 1 0 0.3 - 67.4 0.46 -3.11 12.66
co2_acg_aircraft-pfp_1_allvalid_0-1000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 440 402 26 0 0.1 - 6.8 1.19 -0.59 2.76
co2_acg_aircraft-pfp_1_allvalid_1000-2000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 1490 204 18 0 0.1 - 4.1 1.37 -0.23 1.71
co2_acg_aircraft-pfp_1_allvalid_2000-3000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 2533 118 15 0 0.4 - 2.5 1.71 -0.03 1.26
co2_acg_aircraft-pfp_1_allvalid_3000-4000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 3530 112 11 0 0.2 - 2.1 1.61 0.09 1.28
co2_acg_aircraft-pfp_1_allvalid_4000-5000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 4424 100 7 0 0.4 - 2.0 1.31 -0.19 1.30
co2_acg_aircraft-pfp_1_allvalid_5000-6000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 5494 116 8 0 0.5 - 2.8 1.45 -0.07 1.35
co2_acg_aircraft-pfp_1_allvalid_6000-7000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 6443 117 8 0 0.3 - 3.3 1.69 0.06 1.36
co2_acg_aircraft-pfp_1_allvalid_7000-8000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 7495 149 3 0 0.3 - 2.8 1.38 0.00 1.40
co2_acg_aircraft-pfp_1_allvalid_8000-9000masl NOAA Alaska Coast Guard, United States 57.74°N 152.50°W 8342 20 0 0 0.4 - 1.5 1.32 0.69 0.65
co2_ah2_shipboard-insitu_20_allvalid NIES Alligator Hope (M/S Alligator Hope of Mitsui O.S.K. Lines, Ltd.) variable Surface 27567 2718 0 0.3 - 94.0 1.13 -0.48 2.92
co2_alt_surface-flask_1_representative NOAA Alert, Nunavut, Canada 82.45°N 62.51°W 185 1421 16 0 0.2 - 5.5 0.38 -0.09 0.81
co2_alt_surface-flask_2_representative CSIRO Alert, Nunavut, Canada 82.45°N 62.51°W 185 790 45 0 0.1 - 8.8 1.05 0.04 0.86
co2_alt_surface-flask_4_representative SIO Alert, Nunavut, Canada 82.45°N 62.51°W 185 475 7 0 0.4 - 6.1 0.53 -0.02 0.81
co2_alt_surface-flask_426_representative SIO_CO2 Alert, Nunavut, Canada 82.45°N 62.51°W 185 645 5 0 0.5 - 7.6 0.38 -0.16 0.89
co2_alt_surface-insitu_6_allvalid ECCC Alert, Nunavut, Canada 82.45°N 62.51°W 185 22934 1648 0 0.5 - 3.8 1.16 -0.07 0.79
co2_ams_surface-insitu_11_allvalid LSCE Amsterdam Island, France 37.80°S 77.54°E 55 31170 106 0 0.3 - 1.0 0.41 -0.05 0.28
co2_amt_surface-pfp_1_allvalid-107magl NOAA Argyle, Maine, United States 45.03°N 68.68°W 53 1621 8 0 1.6 - 25.2 0.61 0.16 3.69
co2_amt_tower-insitu_1_allvalid-107magl NOAA Argyle, Maine, United States 45.03°N 68.68°W 53 19889 582 0 1.4 - 8.7 0.98 0.14 3.66
co2_amt_tower-insitu_1_allvalid-12magl NOAA Argyle, Maine, United States 45.03°N 68.68°W 53 18492 565 0 1.5 - 5.7 1.02 0.40 3.97
co2_amt_tower-insitu_1_allvalid-30magl NOAA Argyle, Maine, United States 45.03°N 68.68°W 53 13447 437 0 1.5 - 5.8 1.08 0.64 4.04
co2_ara_surface-flask_2_representative CSIRO Arcturus, Queensland, Australia 23.86°S 148.47°E 175 16 0 0 1.7 - 4.3 0.53 -0.15 2.43
co2_asc_surface-flask_1_representative NOAA Ascension Island, United Kingdom 7.97°S 14.40°W 85 1691 0 0 0.5 - 1.3 1.14 0.13 0.73
co2_ask_surface-flask_1_representative NOAA Assekrem, Algeria 23.26°N 5.63°E 2710 827 22 0 0.3 - 0.9 1.25 -0.23 0.69
co2_azr_surface-flask_1_representative NOAA Terceira Island, Azores, Portugal 38.77°N 27.38°W 19 525 20 0 0.5 - 2.3 1.08 0.07 1.39
co2_azv_tower-insitu_20_allvalid-29magl NIES Azovo, Russia 54.70°N 73.03°E 110 10687 434 0 1.9 - 4.9 1.12 -0.89 3.84
co2_azv_tower-insitu_20_allvalid-50magl NIES Azovo, Russia 54.70°N 73.03°E 110 10387 405 0 1.9 - 5.0 1.11 -0.71 3.82
co2_bal_surface-flask_1_representative NOAA Baltic Sea, Poland 55.35°N 17.22°E 3 899 16 0 0.6 - 12.2 0.82 -2.23 5.54
co2_bao_surface-pfp_1_allvalid-300magl NOAA Boulder Atmospheric Observatory, Colorado, United States 40.05°N 105.00°W 1584 2348 3 0 2.4 - 26.5 0.40 -1.83 2.97
co2_bao_tower-insitu_1_allvalid-100magl NOAA Boulder Atmospheric Observatory, Colorado, United States 40.05°N 105.00°W 1584 9441 209 0 2.2 - 12.9 0.82 -3.10 5.49
co2_bao_tower-insitu_1_allvalid-22magl NOAA Boulder Atmospheric Observatory, Colorado, United States 40.05°N 105.00°W 1584 9555 202 0 2.3 - 16.6 0.80 -3.43 6.38
co2_bao_tower-insitu_1_allvalid-300magl NOAA Boulder Atmospheric Observatory, Colorado, United States 40.05°N 105.00°W 1584 62262 637 0 2.1 - 24.7 0.70 0.14 4.61
co2_bck_surface-insitu_6_allvalid ECCC Behchoko, Northwest Territories, Canada 62.80°N 115.92°W 160 10702 734 0 1.6 - 3.7 1.10 0.06 2.31
co2_bcs_surface-flask_426_representative SIO_CO2 Baja California Sur, Mexico 23.30°N 110.20°W 4 115 1 0 0.3 - 3.8 0.56 0.65 1.83
co2_bgi_aircraft-pfp_1_allvalid_0-1000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 611 5 0 0 2.7 - 7.6 1.04 -2.50 3.07
co2_bgi_aircraft-pfp_1_allvalid_1000-2000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 1587 45 3 0 0.5 - 4.6 1.15 -0.58 2.61
co2_bgi_aircraft-pfp_1_allvalid_2000-3000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 2548 15 1 0 0.5 - 1.6 1.14 0.14 0.92
co2_bgi_aircraft-pfp_1_allvalid_3000-4000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 3524 46 5 0 0.2 - 3.5 1.30 0.19 1.46
co2_bgi_aircraft-pfp_1_allvalid_4000-5000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 4569 23 3 0 0.1 - 1.1 1.58 0.34 0.76
co2_bgi_aircraft-pfp_1_allvalid_5000-6000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 5506 38 0 0 0.3 - 1.5 1.19 0.07 0.97
co2_bgi_aircraft-pfp_1_allvalid_6000-7000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 6474 37 1 0 0.2 - 2.6 1.13 0.19 1.20
co2_bgi_aircraft-pfp_1_allvalid_7000-8000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 7469 35 0 0 0.3 - 1.4 0.77 -0.05 0.69
co2_bgi_aircraft-pfp_1_allvalid_8000-9000masl NOAA Bradgate, Iowa, United States 42.82°N 94.41°W 8050 3 0 0 0.9 - 0.9 0.39 -0.11 1.02
co2_bhd_surface-flask_1_representative NOAA Baring Head Station, New Zealand 41.41°S 174.87°E 85 249 2 0 0.2 - 6.5 0.82 0.04 1.10
co2_bhd_surface-flask_426_representative SIO_CO2 Baring Head Station, New Zealand 41.41°S 174.87°E 85 145 6 0 0.4 - 4.3 0.86 0.99 1.51
co2_bhd_surface-insitu_15_baseline NIWA Baring Head Station, New Zealand 41.41°S 174.87°E 85 673 6 0 0.5 - 4.9 0.65 0.56 0.77
co2_bir_surface-insitu_442_allvalid-10magl ICOS-ATC Birkenes Observatory, Norway 58.39°N 8.25°E 215 207 4 0 2.1 - 3.4 0.97 -0.20 3.15
co2_bir_surface-insitu_56_allvalid NILU Birkenes Observatory, Norway 58.39°N 8.25°E 215 2690 310 0 1.6 - 5.0 1.26 -2.05 5.06
co2_bkt_surface-flask_1_representative NOAA Bukit Kototabang, Indonesia 0.20°S 100.32°E 845 510 0 0 3.8 - 7.5 1.04 4.40 3.74
co2_bme_surface-flask_1_representative NOAA St. Davids Head, Bermuda, United Kingdom 32.37°N 64.65°W 12 221 4 0 0.8 - 2.7 1.02 0.37 1.63
co2_bmw_surface-flask_1_representative NOAA Tudor Hill, Bermuda, United Kingdom 32.26°N 64.88°W 30 675 18 0 0.5 - 2.4 1.09 0.53 1.36
co2_bne_aircraft-pfp_1_allvalid_0-1000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 633 65 2 0 1.5 - 10.6 0.69 -0.65 3.81
co2_bne_aircraft-pfp_1_allvalid_1000-2000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 1383 162 5 0 0.5 - 7.3 1.19 -0.42 2.60
co2_bne_aircraft-pfp_1_allvalid_2000-3000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 2294 106 2 0 0.4 - 8.9 1.00 -0.28 1.86
co2_bne_aircraft-pfp_1_allvalid_3000-4000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 3396 139 6 0 0.1 - 9.7 1.27 -0.19 1.67
co2_bne_aircraft-pfp_1_allvalid_4000-5000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 4278 76 4 0 0.1 - 12.5 1.22 -0.32 2.39
co2_bne_aircraft-pfp_1_allvalid_5000-6000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 5392 109 7 0 0.2 - 14.4 1.14 -0.17 2.20
co2_bne_aircraft-pfp_1_allvalid_6000-7000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 6362 107 3 0 0.2 - 14.9 1.13 -0.13 2.29
co2_bne_aircraft-pfp_1_allvalid_7000-8000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 7657 85 2 0 0.1 - 23.0 1.39 -0.04 3.78
co2_bne_aircraft-pfp_1_allvalid_8000-9000masl NOAA Beaver Crossing, Nebraska, United States 40.80°N 97.18°W 8072 27 0 0 0.4 - 1.2 1.40 -0.49 0.94
co2_bra_surface-insitu_6_allvalid ECCC Bratt's Lake Saskatchewan, Canada 50.20°N 104.71°W 595 10462 437 0 1.8 - 3.9 1.07 -0.06 2.55
co2_brw_surface-flask_1_representative NOAA Barrow Atmospheric Baseline Observatory, United States 71.32°N 156.61°W 11 1834 30 0 0.2 - 13.8 0.57 -0.27 1.43
co2_brw_surface-flask_426_representative SIO_CO2 Barrow Atmospheric Baseline Observatory, United States 71.32°N 156.61°W 11 754 6 0 0.7 - 10.8 0.48 -0.20 1.48
co2_brw_surface-insitu_1_allvalid NOAA Barrow Atmospheric Baseline Observatory, United States 71.32°N 156.61°W 11 34112 1906 0 1.0 - 5.6 0.99 -0.19 1.25
co2_brz_aircraft-insitu_20_allvalid_0-1000masl NIES Berezorechka, Russia 56.15°N 84.33°E 636 30842 190 0 1.5 - 185.6 0.27 -0.38 3.48
co2_brz_aircraft-insitu_20_allvalid_1000-2000masl NIES Berezorechka, Russia 56.15°N 84.33°E 1501 41597 166 0 1.1 - 75.6 0.22 -0.40 2.85
co2_brz_aircraft-insitu_20_allvalid_2000-3000masl NIES Berezorechka, Russia 56.15°N 84.33°E 2408 18582 61 0 0.9 - 64.1 0.21 -0.27 2.26
co2_brz_aircraft-insitu_20_allvalid_3000-4000masl NIES Berezorechka, Russia 56.15°N 84.33°E 3085 2446 24 0 0.4 - 41.2 0.27 -0.31 2.17
co2_brz_tower-insitu_20_allvalid-20magl NIES Berezorechka, Russia 56.15°N 84.33°E 168 13640 563 0 2.1 - 19.0 1.03 -0.16 3.94
co2_brz_tower-insitu_20_allvalid-40magl NIES Berezorechka, Russia 56.15°N 84.33°E 168 13167 515 0 2.1 - 18.5 1.04 -0.42 3.94
co2_brz_tower-insitu_20_allvalid-5magl NIES Berezorechka, Russia 56.15°N 84.33°E 168 13780 520 0 2.2 - 18.9 1.04 -0.22 4.01
co2_brz_tower-insitu_20_allvalid-80magl NIES Berezorechka, Russia 56.15°N 84.33°E 168 9082 326 0 1.9 - 19.0 1.06 -0.69 4.04
co2_bsc_surface-flask_1_representative NOAA Black Sea, Constanta, Romania 44.18°N 28.66°E 0 409 3 0 2.3 - 16.3 0.94 -5.69 7.99
co2_car_aircraft-pfp_1_allvalid_1000-2000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 1783 48 1 0 1.2 - 4.2 0.67 -0.27 2.23
co2_car_aircraft-pfp_1_allvalid_2000-3000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 2450 969 20 0 0.3 - 5.1 0.84 0.20 1.90
co2_car_aircraft-pfp_1_allvalid_3000-4000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 3455 1124 47 0 0.2 - 2.3 1.18 0.10 0.98
co2_car_aircraft-pfp_1_allvalid_4000-5000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 4497 1024 37 0 0.2 - 2.1 1.16 0.21 0.85
co2_car_aircraft-pfp_1_allvalid_5000-6000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 5485 785 26 0 0.3 - 2.1 1.12 0.18 0.80
co2_car_aircraft-pfp_1_allvalid_6000-7000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 6438 773 32 0 0.5 - 2.0 1.19 0.35 0.76
co2_car_aircraft-pfp_1_allvalid_7000-8000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 7473 764 9 0 0.3 - 2.2 1.16 0.29 0.77
co2_car_aircraft-pfp_1_allvalid_8000-9000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 8219 170 2 0 0.2 - 1.7 1.16 0.12 0.76
co2_car_aircraft-pfp_1_allvalid_9000-10000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 9140 1 0 0 0.7 - 0.7 1.52 0.46 NA
co2_car_aircraft-pfp_1_allvalid_11000-12000masl NOAA Briggsdale, Colorado, United States 40.63°N 104.33°W 11869 2 0 0 0.5 - 0.5 1.47 0.22 0.60
co2_cba_surface-flask_1_representative NOAA Cold Bay, Alaska, United States 55.21°N 162.72°W 21 1404 29 0 0.8 - 4.8 1.01 -0.93 1.68
co2_cba_surface-flask_4_representative SIO Cold Bay, Alaska, United States 55.21°N 162.72°W 21 399 15 0 0.6 - 6.5 1.16 -0.70 2.01
co2_cby_surface-insitu_6_allvalid ECCC Cambridge Bay, Nunavut Territory, Canada 69.13°N 105.06°W 35 8767 691 0 0.8 - 1.8 1.15 -0.03 1.22
co2_cdl_surface-insitu_6_allvalid ECCC Candle Lake, Saskatchewan, Canada 53.99°N 105.12°W 600 8503 316 0 1.6 - 4.3 0.93 0.09 2.54
co2_cfa_surface-flask_2_representative CSIRO Cape Ferguson, Queensland, Australia 19.28°S 147.06°E 2 583 0 0 0.1 - 2.4 0.51 -0.35 1.06
co2_cgo_surface-flask_1_representative NOAA Cape Grim, Tasmania, Australia 40.68°S 144.69°E 94 703 0 0 0.3 - 7.3 0.74 -0.18 0.64
co2_cgo_surface-flask_2_representative CSIRO Cape Grim, Tasmania, Australia 40.68°S 144.69°E 94 1228 4 0 0.3 - 4.0 0.47 -0.25 0.61
co2_cgo_surface-flask_4_representative SIO Cape Grim, Tasmania, Australia 40.68°S 144.69°E 94 411 7 0 0.3 - 1.2 0.86 -0.29 0.56
co2_chl_surface-insitu_6_allvalid ECCC Churchill, Manitoba, Canada 58.74°N 93.82°W 29 7617 695 0 1.1 - 3.0 1.04 -0.14 1.82
co2_chm_surface-insitu_6_allvalid ECCC Chibougamau, Quebec, Canada 49.69°N 74.34°W 393 3400 172 0 1.9 - 3.5 1.15 0.05 2.48
co2_chr_surface-flask_1_representative NOAA Christmas Island, Republic of Kiribati 1.70°N 157.15°W 0 594 0 0 0.3 - 1.8 0.67 -0.20 0.60
co2_chr_surface-flask_426_representative SIO_CO2 Christmas Island, Republic of Kiribati 1.70°N 157.15°W 0 273 2 0 0.5 - 2.2 0.83 -0.50 0.77
co2_cib_surface-flask_1_representative NOAA Centro de Investigacion de la Baja Atmosfera (CIBA), Spain 41.81°N 4.93°W 845 433 10 0 1.9 - 5.0 1.10 0.82 3.83
co2_cma_aircraft-pfp_1_allvalid_0-1000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 635 511 11 0 1.5 - 7.6 1.14 -0.29 3.09
co2_cma_aircraft-pfp_1_allvalid_1000-2000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 1536 311 6 0 0.7 - 5.0 0.98 -0.31 2.24
co2_cma_aircraft-pfp_1_allvalid_2000-3000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 2301 392 5 0 0.6 - 6.1 0.92 -0.19 1.80
co2_cma_aircraft-pfp_1_allvalid_3000-4000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 3441 358 13 0 0.3 - 2.8 1.10 0.18 1.21
co2_cma_aircraft-pfp_1_allvalid_4000-5000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 4190 189 7 0 0.3 - 2.7 1.20 -0.03 1.31
co2_cma_aircraft-pfp_1_allvalid_5000-6000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 5345 350 10 0 0.4 - 2.3 1.16 0.09 1.10
co2_cma_aircraft-pfp_1_allvalid_6000-7000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 6269 287 7 0 0.4 - 1.8 1.14 -0.09 1.02
co2_cma_aircraft-pfp_1_allvalid_7000-8000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 7725 274 4 0 0.4 - 1.8 1.10 0.28 0.88
co2_cma_aircraft-pfp_1_allvalid_8000-9000masl NOAA Offshore Cape May, New Jersey, United States 38.83°N 74.32°W 8048 33 0 0 0.6 - 1.4 0.76 -0.06 0.98
co2_con_aircraft-flask_42_allvalid_0-1000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 820 1 0 0 0.1 - 0.1 0.78 -0.09 NA
co2_con_aircraft-flask_42_allvalid_1000-2000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 1385 4 0 0 0.3 - 0.4 0.57 0.23 0.35
co2_con_aircraft-flask_42_allvalid_3000-4000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 3491 2 0 0 0.3 - 0.3 2.53 -0.37 0.34
co2_con_aircraft-flask_42_allvalid_5000-6000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 5631 2 0 0 1.0 - 1.0 0.44 -0.02 0.83
co2_con_aircraft-flask_42_allvalid_6000-7000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 6391 4 0 0 0.1 - 2.7 1.56 1.14 1.64
co2_con_aircraft-flask_42_allvalid_8000-9000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 8578 6 0 0 0.1 - 0.8 1.07 -0.06 0.46
co2_con_aircraft-flask_42_allvalid_9000-10000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 9648 305 5 1 0.0 - 2.2 1.18 0.13 0.71
co2_con_aircraft-flask_42_allvalid_10000-11000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 10672 1923 36 0 0.1 - 1.6 0.98 0.06 0.64
co2_con_aircraft-flask_42_allvalid_11000-12000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 11536 1023 16 0 0.1 - 1.7 0.89 0.13 0.72
co2_con_aircraft-flask_42_allvalid_12000-13000masl NIES CONTRAIL (Comprehensive Observation Network for TRace gases by AIrLiner) variable 12260 208 9 0 0.1 - 2.2 0.95 -0.08 0.97
co2_cps_surface-insitu_6_allvalid ECCC Chapais,Quebec, Canada 49.82°N 74.98°W 391 9312 707 0 1.4 - 3.6 1.41 0.25 2.73
co2_cpt_surface-flask_1_representative NOAA Cape Point, South Africa 34.35°S 18.49°E 230 221 2 0 0.2 - 1.3 0.44 0.18 0.43
co2_cpt_surface-insitu_36_marine SAWS Cape Point, South Africa 34.35°S 18.49°E 230 133019 5268 0 0.4 - 1.0 1.00 0.00 0.56
co2_crv_aircraft-pfp_1_allvalid_0-1000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 315 1390 50 0 0.9 - 50.8 0.93 -2.28 5.51
co2_crv_aircraft-pfp_1_allvalid_1000-2000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 1438 60 19 0 0.3 - 8.8 1.72 0.45 2.39
co2_crv_aircraft-pfp_1_allvalid_2000-3000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 2555 72 2 0 0.2 - 3.3 1.20 0.12 1.46
co2_crv_aircraft-pfp_1_allvalid_3000-4000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 3390 48 6 0 0.4 - 1.9 1.98 0.25 1.09
co2_crv_aircraft-pfp_1_allvalid_4000-5000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 4518 27 4 0 0.2 - 1.9 2.33 0.51 1.12
co2_crv_aircraft-pfp_1_allvalid_5000-6000masl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 5269 267 14 0 0.1 - 2.2 1.12 0.20 1.42
co2_crv_surface-pfp_1_allvalid-32magl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 611 1332 10 0 1.9 - 9.2 0.66 -0.30 3.16
co2_crv_tower-insitu_1_allvalid-17magl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 611 9284 560 0 1.5 - 4.8 1.23 -0.19 3.04
co2_crv_tower-insitu_1_allvalid-32magl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 611 9590 700 0 1.5 - 5.6 1.26 -0.26 2.83
co2_crv_tower-insitu_1_allvalid-5magl NOAA Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE), United States 64.99°N 147.60°W 611 9357 678 0 1.5 - 4.3 1.30 -0.10 3.16
co2_crz_surface-flask_1_representative NOAA Crozet Island, France 46.43°S 51.85°E 197 756 0 0 0.1 - 0.5 0.82 0.01 0.32
co2_cya_surface-flask_2_representative CSIRO Casey, Antarctica, Australia 66.28°S 110.52°E 47 576 0 0 0.1 - 0.6 0.81 0.03 0.25
co2_dem_tower-insitu_20_allvalid-45magl NIES Demyanskoe, Russia 59.79°N 70.87°E 63 13827 497 0 1.7 - 5.0 1.19 -0.33 3.34
co2_dem_tower-insitu_20_allvalid-63magl NIES Demyanskoe, Russia 59.79°N 70.87°E 63 12988 505 0 1.7 - 5.1 1.20 -0.45 3.36
co2_dnd_aircraft-pfp_1_allvalid_0-1000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 744 212 8 0 0.4 - 7.0 1.04 -0.49 3.69
co2_dnd_aircraft-pfp_1_allvalid_1000-2000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 1507 281 24 0 0.2 - 5.5 1.10 -0.23 2.62
co2_dnd_aircraft-pfp_1_allvalid_2000-3000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 2470 312 31 0 0.1 - 3.2 1.27 -0.18 1.42
co2_dnd_aircraft-pfp_1_allvalid_3000-4000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 3530 248 14 0 0.4 - 2.9 1.57 0.03 1.27
co2_dnd_aircraft-pfp_1_allvalid_4000-5000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 4466 187 7 0 0.5 - 3.8 1.14 0.07 1.31
co2_dnd_aircraft-pfp_1_allvalid_5000-6000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 5508 216 9 0 0.3 - 2.4 1.15 0.13 1.05
co2_dnd_aircraft-pfp_1_allvalid_6000-7000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 6481 196 4 0 0.3 - 2.3 1.37 0.11 1.04
co2_dnd_aircraft-pfp_1_allvalid_7000-8000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 7474 201 8 0 0.2 - 1.9 1.28 0.22 0.99
co2_dnd_aircraft-pfp_1_allvalid_8000-9000masl NOAA Dahlen, North Dakota, United States 47.50°N 99.24°W 8053 3 0 0 0.6 - 2.3 1.06 -1.37 1.02
co2_drp_shipboard-flask_1_representative NOAA Drake Passage variable Surface 243 5 0 0.1 - 1.2 0.65 0.03 0.40
co2_dsi_surface-flask_1_representative NOAA Dongsha Island, Taiwan 20.70°N 116.73°E 3 385 0 0 1.9 - 6.9 0.82 1.19 2.81
co2_egb_surface-insitu_6_allvalid ECCC Egbert, Ontario, Canada 44.23°N 79.78°W 251 13640 426 0 2.5 - 4.9 0.95 -0.21 3.71
co2_eic_surface-flask_1_representative NOAA Easter Island, Chile 27.16°S 109.43°W 47 540 2 0 0.5 - 1.7 1.02 0.36 0.97
co2_esp_aircraft-pfp_1_allvalid_0-1000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 544 744 35 0 0.6 - 7.3 0.79 -0.76 3.02
co2_esp_aircraft-pfp_1_allvalid_1000-2000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 1554 931 56 0 0.3 - 3.2 1.26 -0.11 1.40
co2_esp_aircraft-pfp_1_allvalid_2000-3000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 2554 800 44 0 0.2 - 2.5 1.23 -0.05 1.22
co2_esp_aircraft-pfp_1_allvalid_3000-4000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 3553 757 41 0 0.3 - 2.4 1.36 0.05 1.15
co2_esp_aircraft-pfp_1_allvalid_4000-5000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 4511 662 22 0 0.5 - 2.4 1.26 0.08 1.16
co2_esp_aircraft-pfp_1_allvalid_5000-6000masl NOAA Estevan Point, British Columbia, Canada 49.38°N 126.54°W 5378 446 10 0 0.5 - 2.0 1.16 -0.00 1.10
co2_esp_surface-flask_2_representative CSIRO Estevan Point, British Columbia, Canada 49.38°N 126.54°W 7 17 2 0 0.2 - 7.1 0.48 -0.97 1.71
co2_esp_surface-insitu_6_allvalid ECCC Estevan Point, British Columbia, Canada 49.38°N 126.54°W 7 12971 393 0 1.6 - 3.8 0.84 -0.41 2.17
co2_est_surface-insitu_6_allvalid ECCC Esther, Alberta, Canada 51.67°N 110.21°W 707 11753 415 0 1.8 - 4.4 1.09 -0.19 2.78
co2_etl_aircraft-pfp_1_allvalid_0-1000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 909 194 18 0 1.0 - 3.8 1.30 -0.34 2.23
co2_etl_aircraft-pfp_1_allvalid_1000-2000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 1511 741 40 0 0.7 - 3.1 1.26 -0.25 1.97
co2_etl_aircraft-pfp_1_allvalid_2000-3000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 2484 789 28 0 0.8 - 5.5 1.19 -0.26 1.49
co2_etl_aircraft-pfp_1_allvalid_3000-4000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 3501 279 11 0 0.5 - 2.1 1.23 -0.03 1.23
co2_etl_aircraft-pfp_1_allvalid_4000-5000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 4580 269 10 0 0.6 - 5.1 1.00 0.05 1.50
co2_etl_aircraft-pfp_1_allvalid_5000-6000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 5642 254 10 0 0.6 - 2.1 1.25 0.05 1.21
co2_etl_aircraft-pfp_1_allvalid_6000-7000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 6741 136 3 0 0.5 - 1.9 1.46 0.48 1.02
co2_etl_aircraft-pfp_1_allvalid_7000-8000masl NOAA East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 7150 96 3 0 0.7 - 2.4 1.26 0.02 1.71
co2_etl_surface-insitu_6_allvalid ECCC East Trout Lake, Saskatchewan, Canada 54.35°N 104.99°W 493 17218 1007 0 1.6 - 3.9 1.07 -0.14 2.45
co2_fsd_surface-insitu_6_allvalid ECCC Fraserdale, Canada 49.88°N 81.57°W 210 21843 999 0 1.7 - 4.5 1.05 0.03 2.68
co2_ftl_aircraft-pfp_1_allvalid_1000-2000masl NOAA Fortaleza, Brazil 3.52°S 38.28°W 1810 10 0 0 0.1 - 1.2 1.42 -0.46 1.09
co2_ftl_aircraft-pfp_1_allvalid_2000-3000masl NOAA Fortaleza, Brazil 3.52°S 38.28°W 2498 23 0 0 0.4 - 2.2 0.97 -0.53 1.28
co2_ftl_aircraft-pfp_1_allvalid_3000-4000masl NOAA Fortaleza, Brazil 3.52°S 38.28°W 3479 37 0 0 0.3 - 2.1 1.24 0.04 1.39
co2_ftl_aircraft-pfp_1_allvalid_4000-5000masl NOAA Fortaleza, Brazil 3.52°S 38.28°W 4267 7 0 0 0.5 - 2.7 1.12 0.07 1.64
co2_ftw_shipboard-insitu_20_allvalid NIES Fujitrans World (M/S Fujitrans World of Kagoshima Shipping Co., Ltd.) variable Surface 69451 2688 0 0.1 - 19.6 0.85 -0.24 1.51
co2_ftws_shipboard-insitu_20_allvalid NIES Fujitrans World - Southeast Asia Route (M/S Fujitrans World of Kagoshima Shipping Co., Ltd.) variable Surface 227646 6095 0 0.1 - 32.5 0.69 -0.86 2.83
co2_fwi_aircraft-pfp_1_allvalid_0-1000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 623 9 0 0 3.6 - 10.7 0.54 -1.50 5.32
co2_fwi_aircraft-pfp_1_allvalid_1000-2000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 1577 47 4 0 0.5 - 8.2 1.22 -0.54 2.78
co2_fwi_aircraft-pfp_1_allvalid_2000-3000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 2535 18 0 0 0.3 - 2.8 1.02 -0.37 1.95
co2_fwi_aircraft-pfp_1_allvalid_3000-4000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 3518 48 7 0 0.2 - 2.9 1.32 0.04 1.68
co2_fwi_aircraft-pfp_1_allvalid_4000-5000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 4570 28 2 0 0.2 - 2.1 1.30 0.32 1.03
co2_fwi_aircraft-pfp_1_allvalid_5000-6000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 5523 41 1 0 0.2 - 9.9 1.08 0.72 2.37
co2_fwi_aircraft-pfp_1_allvalid_6000-7000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 6504 30 5 0 0.1 - 2.2 1.65 0.37 1.11
co2_fwi_aircraft-pfp_1_allvalid_7000-8000masl NOAA Fairchild, Wisconsin, United States 44.66°N 90.96°W 7484 39 0 0 0.5 - 2.2 1.26 0.13 1.15
co2_gmi_surface-flask_1_representative NOAA Mariana Islands, Guam 13.39°N 144.66°E 0 1072 21 0 0.3 - 1.4 0.89 0.07 0.78
co2_gw_shipboard-insitu_20_allvalid NIES Golden Wattle (M/S Alligator Hope of Mitsui O.S.K. Lines, Ltd.) variable Surface 13101 877 0 0.1 - 58.7 0.96 -0.27 1.56
co2_haa_aircraft-pfp_1_allvalid_0-1000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 879 20 2 0 0.2 - 1.2 1.75 0.36 0.63
co2_haa_aircraft-pfp_1_allvalid_1000-2000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 1598 206 14 0 0.2 - 1.0 1.28 0.17 0.70
co2_haa_aircraft-pfp_1_allvalid_2000-3000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 2524 196 15 0 0.2 - 1.1 1.35 0.08 0.73
co2_haa_aircraft-pfp_1_allvalid_3000-4000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 3488 217 8 0 0.2 - 1.1 1.21 0.07 0.71
co2_haa_aircraft-pfp_1_allvalid_4000-5000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 4532 251 7 0 0.1 - 1.2 1.27 0.10 0.80
co2_haa_aircraft-pfp_1_allvalid_5000-6000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 5440 194 3 0 0.1 - 1.1 1.26 0.07 0.75
co2_haa_aircraft-pfp_1_allvalid_6000-7000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 6490 202 1 0 0.2 - 1.5 1.23 0.21 0.94
co2_haa_aircraft-pfp_1_allvalid_7000-8000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 7471 114 2 0 0.2 - 2.1 0.91 0.24 1.30
co2_haa_aircraft-pfp_1_allvalid_8000-9000masl NOAA Molokai Island, Hawaii, United States 21.23°N 158.95°W 8041 61 0 0 0.2 - 3.1 1.06 -0.10 0.86
co2_hba_surface-flask_1_representative NOAA Halley Station, Antarctica, United Kingdom 75.61°S 26.21°W 30 726 0 0 0.0 - 0.6 1.01 0.05 0.18
co2_hdp_surface-insitu_3_nonlocal NCAR Hidden Peak (Snowbird), Utah, United States 40.56°N 111.65°W 3351 55241 1975 0 0.8 - 2.0 1.18 -0.36 1.19
co2_hfm_aircraft-pfp_1_allvalid_0-1000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 761 146 8 0 0.3 - 7.0 0.76 -0.40 3.46
co2_hfm_aircraft-pfp_1_allvalid_1000-2000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 1532 253 10 0 1.0 - 10.7 0.97 -0.20 2.53
co2_hfm_aircraft-pfp_1_allvalid_2000-3000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 2452 174 13 0 0.5 - 8.9 1.46 -0.19 2.00
co2_hfm_aircraft-pfp_1_allvalid_3000-4000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 3432 149 8 0 0.3 - 6.2 1.29 0.11 1.27
co2_hfm_aircraft-pfp_1_allvalid_4000-5000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 4563 181 3 0 0.3 - 2.5 1.34 0.17 1.03
co2_hfm_aircraft-pfp_1_allvalid_5000-6000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 5483 196 4 0 0.3 - 1.7 1.26 0.27 0.95
co2_hfm_aircraft-pfp_1_allvalid_6000-7000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 6426 145 3 0 0.3 - 2.5 1.62 0.42 0.94
co2_hfm_aircraft-pfp_1_allvalid_7000-8000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 7388 167 7 0 0.2 - 6.5 1.14 0.44 2.33
co2_hfm_aircraft-pfp_1_allvalid_8000-9000masl NOAA Harvard Forest, Massachusetts, United States 42.54°N 72.17°W 8031 2 0 0 1.7 - 1.7 0.31 -1.11 1.31
co2_hil_aircraft-pfp_1_allvalid_0-1000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 612 174 6 0 0.9 - 8.0 1.13 -0.67 3.40
co2_hil_aircraft-pfp_1_allvalid_1000-2000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 1554 436 16 0 0.7 - 6.3 1.05 -0.25 2.58
co2_hil_aircraft-pfp_1_allvalid_2000-3000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 2537 255 13 0 0.1 - 4.8 0.90 -0.45 1.77
co2_hil_aircraft-pfp_1_allvalid_3000-4000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 3503 454 15 0 0.3 - 3.0 1.05 -0.27 1.42
co2_hil_aircraft-pfp_1_allvalid_4000-5000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 4545 295 9 0 0.2 - 2.5 1.25 -0.13 1.22
co2_hil_aircraft-pfp_1_allvalid_5000-6000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 5513 389 7 0 0.2 - 2.0 1.20 -0.15 1.20
co2_hil_aircraft-pfp_1_allvalid_6000-7000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 6520 323 22 0 0.2 - 2.3 1.15 0.02 1.19
co2_hil_aircraft-pfp_1_allvalid_7000-8000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 7484 357 10 0 0.4 - 2.5 1.06 -0.04 1.10
co2_hil_aircraft-pfp_1_allvalid_8000-9000masl NOAA Homer, Illinois, United States 40.07°N 87.91°W 8042 32 0 0 0.7 - 3.1 1.20 -1.15 1.78
co2_hpb_surface-flask_1_representative NOAA Hohenpeissenberg, Germany 47.80°N 11.02°E 936 600 12 0 3.6 - 14.5 1.08 2.03 6.81
co2_hun_surface-flask_1_representative NOAA Hegyhatsal, Hungary 46.95°N 16.65°E 248 937 10 0 2.1 - 9.9 0.79 -2.01 5.47
co2_ice_surface-flask_1_representative NOAA Storhofdi, Vestmannaeyjar, Iceland 63.40°N 20.29°W 118 538 23 0 0.4 - 4.2 1.03 -0.45 1.07
co2_igr_tower-insitu_20_allvalid-24magl NIES Igrim, Russia 63.19°N 64.42°E 9 10627 304 0 3.7 - 6.0 0.92 -1.96 4.43
co2_igr_tower-insitu_20_allvalid-47magl NIES Igrim, Russia 63.19°N 64.42°E 9 10515 274 0 3.9 - 8.4 0.68 -1.86 5.91
co2_inu_surface-insitu_6_allvalid ECCC Inuvik,Northwest Territories, Canada 68.32°N 133.53°W 113 10857 591 0 1.9 - 3.7 1.05 -0.26 2.31
co2_inx_aircraft-pfp_1_allvalid_0-1000masl NOAA INFLUX (Indianapolis Flux Experiment), United States 39.58°N 86.42°W 652 171 3 0 1.1 - 5.8 0.81 -2.14 3.81
co2_inx_aircraft-pfp_1_allvalid_1000-2000masl NOAA INFLUX (Indianapolis Flux Experiment), United States 39.58°N 86.42°W 1354 57 3 0 0.8 - 5.2 1.23 -0.49 2.62
co2_inx_aircraft-pfp_1_allvalid_2000-3000masl NOAA INFLUX (Indianapolis Flux Experiment), United States 39.58°N 86.42°W 2501 21 0 0 0.2 - 2.1 0.92 0.26 1.53
co2_inx_aircraft-pfp_1_allvalid_3000-4000masl NOAA INFLUX (Indianapolis Flux Experiment), United States 39.58°N 86.42°W 3226 9 1 0 0.2 - 1.0 2.72 0.24 0.67
co2_izo_surface-insitu_27_allvalid AEMET Izana, Tenerife, Canary Islands, Spain 28.31°N 16.50°W 2373 55987 1830 14813 0.6 - 1.0 1.08 0.18 0.75
co2_jfj_surface-insitu_442_allvalid-5magl ICOS-ATC Jungfraujoch, Switzerland 46.55°N 7.99°E 3570 5142 22 0 0.9 - 3.9 0.44 0.15 1.50
co2_jfj_surface-insitu_49_allvalid KUP Jungfraujoch, Switzerland 46.55°N 7.99°E 3570 19147 331 0 1.5 - 4.5 0.66 0.09 1.85
co2_jfj_surface-insitu_5_allvalid EMPA Jungfraujoch, Switzerland 46.55°N 7.99°E 3570 7744 109 0 0.8 - 5.3 0.61 0.17 1.68
co2_key_surface-flask_1_representative NOAA Key Biscayne, Florida, United States 25.67°N 80.16°W 1 649 13 0 1.0 - 2.5 0.64 0.21 1.75
co2_krs_tower-insitu_20_allvalid-35magl NIES Karasevoe, Russia 58.25°N 82.42°E 76 13234 615 0 1.8 - 5.8 1.14 -0.39 3.77
co2_krs_tower-insitu_20_allvalid-67magl NIES Karasevoe, Russia 58.25°N 82.42°E 76 12452 611 0 1.8 - 5.7 1.16 -0.45 3.69
co2_kum_surface-flask_1_representative NOAA Cape Kumukahi, Hawaii, United States 19.56°N 154.89°W 8 2098 14 0 0.3 - 3.4 0.58 -0.21 0.95
co2_kum_surface-flask_4_representative SIO Cape Kumukahi, Hawaii, United States 19.56°N 154.89°W 8 664 25 0 0.4 - 1.3 0.89 -0.18 1.04
co2_kum_surface-flask_426_representative SIO_CO2 Cape Kumukahi, Hawaii, United States 19.56°N 154.89°W 8 699 15 0 0.3 - 1.6 0.94 -0.17 0.91
co2_kzd_surface-flask_1_representative NOAA Sary Taukum, Kazakhstan 44.08°N 76.87°E 595 414 6 0 0.5 - 8.6 0.74 -1.16 3.78
co2_kzm_surface-flask_1_representative NOAA Plateau Assy, Kazakhstan 43.25°N 77.88°E 2519 361 5 0 1.2 - 4.1 0.96 -0.09 2.61
co2_lef_aircraft-pfp_1_allvalid_0-1000masl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 773 618 39 0 0.9 - 6.7 1.00 -0.39 2.97
co2_lef_aircraft-pfp_1_allvalid_1000-2000masl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 1522 1196 50 0 0.4 - 6.3 0.99 -0.21 2.65
co2_lef_aircraft-pfp_1_allvalid_2000-3000masl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 2464 939 35 0 0.5 - 4.2 0.88 -0.27 1.80
co2_lef_aircraft-pfp_1_allvalid_3000-4000masl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 3502 1085 18 0 0.3 - 3.3 0.96 -0.19 1.48
co2_lef_aircraft-pfp_1_allvalid_4000-5000masl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 4016 3 0 0 1.8 - 1.8 0.43 -0.91 1.33
co2_lef_tower-insitu_1_allvalid-11magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 10478 370 0 2.1 - 4.7 0.94 -0.07 3.41
co2_lef_tower-insitu_1_allvalid-122magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 25084 1266 0 1.9 - 5.0 1.02 -0.08 3.56
co2_lef_tower-insitu_1_allvalid-244magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 62630 2346 0 2.0 - 7.4 0.94 -0.25 3.59
co2_lef_tower-insitu_1_allvalid-30magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 24555 1190 0 2.0 - 5.3 1.02 0.04 3.73
co2_lef_tower-insitu_1_allvalid-396magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 150691 6560 0 1.9 - 13.8 0.99 -0.05 3.41
co2_lef_tower-insitu_1_allvalid-76magl NOAA Park Falls, Wisconsin, United States 45.95°N 90.27°W 472 10330 450 0 2.0 - 4.4 0.96 -0.34 3.30
co2_lew_surface-pfp_1_allvalid-95magl NOAA Lewisburg, Pennsylvania, United States 40.94°N 76.88°W 166 688 19 0 3.1 - 9.4 0.74 -1.73 5.53
co2_ljo_surface-flask_426_representative SIO_CO2 La Jolla, California, United States 32.87°N 117.26°W 10 448 1 0 1.0 - 16.0 0.40 1.91 2.61
co2_llb_surface-flask_1_representative NOAA Lac La Biche, Alberta, Canada 54.95°N 112.47°W 540 142 3 0 2.0 - 18.0 0.57 -1.44 4.66
co2_llb_surface-insitu_6_allvalid ECCC Lac La Biche, Alberta, Canada 54.95°N 112.47°W 540 13413 384 0 2.0 - 6.5 1.00 -0.61 3.44
co2_lmp_surface-flask_1_representative NOAA Lampedusa, Italy 35.52°N 12.63°E 45 559 10 0 0.9 - 2.5 1.07 -0.12 1.73
co2_lut_surface-insitu_44_allvalid RUG Lutjewad, Netherlands 53.40°N 6.35°E 1 12528 389 0 3.5 - 11.4 0.75 -2.18 6.07
co2_maa_surface-flask_2_representative CSIRO Mawson Station, Antarctica, Australia 67.62°S 62.87°E 32 664 0 0 0.1 - 1.2 0.76 0.03 0.23
co2_mbo_surface-pfp_1_allvalid-11magl NOAA Mt. Bachelor Observatory, United States 43.98°N 121.69°W 2731 1465 23 0 0.9 - 4.6 0.72 -0.24 1.62
co2_mex_surface-flask_1_representative NOAA High Altitude Global Climate Observation Center, Mexico 18.98°N 97.31°W 4464 359 4 0 0.8 - 3.2 0.97 0.80 1.40
co2_mhd_surface-flask_1_representative NOAA Mace Head, County Galway, Ireland 53.33°N 9.90°W 5 787 16 0 0.6 - 4.6 0.80 -0.34 1.16
co2_mhd_surface-insitu_11_representative LSCE Mace Head, County Galway, Ireland 53.33°N 9.90°W 5 34349 2401 0 0.6 - 2.2 1.37 -0.06 0.90
co2_mid_surface-flask_1_representative NOAA Sand Island, Midway, United States 28.21°N 177.38°W 11 868 20 0 0.3 - 1.5 1.13 0.34 1.00
co2_mkn_surface-flask_1_representative NOAA Mt. Kenya, Kenya 0.06°S 37.30°E 3644 127 0 0 0.4 - 3.3 1.10 1.66 1.89
co2_mlo_surface-flask_1_representative NOAA Mauna Loa, Hawaii, United States 19.54°N 155.58°W 3397 1962 30 0 0.4 - 2.6 0.61 0.09 0.53
co2_mlo_surface-flask_2_representative CSIRO Mauna Loa, Hawaii, United States 19.54°N 155.58°W 3397 976 1 0 0.5 - 3.3 0.25 0.24 0.56
co2_mlo_surface-flask_4_representative SIO Mauna Loa, Hawaii, United States 19.54°N 155.58°W 3397 761 13 0 0.3 - 1.4 0.76 0.12 0.57
co2_mlo_surface-flask_426_representative SIO_CO2 Mauna Loa, Hawaii, United States 19.54°N 155.58°W 3397 956 30 0 0.4 - 1.4 1.02 0.08 0.54
co2_mlo_surface-insitu_1_allvalid NOAA Mauna Loa, Hawaii, United States 19.54°N 155.58°W 3397 38188 0 0 0.4 - 0.7 1.34 0.14 0.52
co2_mqa_surface-flask_2_representative CSIRO Macquarie Island, Australia 54.48°S 158.97°E 6 733 0 0 0.2 - 1.0 0.58 0.21 0.36
co2_mvy_surface-insitu_1_allvalid NOAA Marthas Vineyard, Massachusetts, United States 41.33°N 70.57°W 0 63052 0 0 2.5 - 8.2 1.10 -0.21 4.28
co2_mwo_surface-pfp_1_allvalid-46magl NOAA Mt. Wilson Observatory, United States 34.22°N 118.06°W 1728 3926 76 0 1.3 - 17.3 0.68 -2.80 5.35
co2_nat_surface-flask_1_representative NOAA Farol De Mae Luiza Lighthouse, Brazil 5.80°S 35.19°W 50 310 3 0 0.8 - 1.6 1.03 -0.68 1.02
co2_nat_surface-flask_26_marine IPEN Farol De Mae Luiza Lighthouse, Brazil 5.80°S 35.19°W 50 194 1 0 0.7 - 2.0 0.95 -0.81 1.13
co2_nha_aircraft-pfp_1_allvalid_0-1000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 612 762 22 0 1.4 - 6.6 1.11 -0.16 2.72
co2_nha_aircraft-pfp_1_allvalid_1000-2000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 1494 595 21 0 1.0 - 4.0 1.06 -0.08 2.10
co2_nha_aircraft-pfp_1_allvalid_2000-3000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 2407 576 25 0 0.3 - 3.7 1.00 -0.21 1.79
co2_nha_aircraft-pfp_1_allvalid_3000-4000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 3482 480 24 0 0.4 - 3.6 1.17 0.04 1.33
co2_nha_aircraft-pfp_1_allvalid_4000-5000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 4410 285 11 0 0.4 - 2.1 1.23 -0.09 1.19
co2_nha_aircraft-pfp_1_allvalid_5000-6000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 5368 360 15 0 0.3 - 3.5 1.10 -0.01 1.26
co2_nha_aircraft-pfp_1_allvalid_6000-7000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 6304 268 6 0 0.3 - 3.6 1.21 -0.05 1.27
co2_nha_aircraft-pfp_1_allvalid_7000-8000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 7666 287 6 0 0.5 - 2.0 1.31 0.28 1.11
co2_nha_aircraft-pfp_1_allvalid_8000-9000masl NOAA Offshore Portsmouth, New Hampshire (Isles of Shoals), United States 42.95°N 70.63°W 8039 4 1 0 0.4 - 0.5 1.01 -0.60 0.71
co2_nmb_surface-flask_1_representative NOAA Gobabeb, Namibia 23.58°S 15.03°E 456 485 14 0 0.5 - 1.8 0.94 0.05 0.98
co2_noy_tower-insitu_20_allvalid-21magl NIES Noyabrsk, Russia 63.43°N 75.78°E 108 9349 414 0 1.5 - 5.2 1.12 -0.01 3.15
co2_noy_tower-insitu_20_allvalid-43magl NIES Noyabrsk, Russia 63.43°N 75.78°E 108 9714 426 0 1.6 - 5.4 1.13 -0.14 3.13
co2_nwr_surface-flask_1_representative NOAA Niwot Ridge, Colorado, United States 40.05°N 105.59°W 3523 1265 4 0 0.5 - 8.3 0.53 0.37 1.29
co2_nwr_surface-insitu_3_nonlocal NCAR Niwot Ridge, Colorado, United States 40.05°N 105.59°W 3523 62986 2084 0 0.8 - 4.5 0.97 -0.04 1.40
co2_nwr_surface-pfp_1_allvalid-3magl NOAA Niwot Ridge, Colorado, United States 40.05°N 105.59°W 3523 3212 56 0 0.7 - 10.5 0.52 0.13 1.81
co2_obn_surface-flask_1_representative NOAA Obninsk, Russia 55.11°N 36.60°E 183 173 3 0 2.6 - 12.1 0.56 -0.18 5.36
co2_ofr_surface-insitu_68_allhours OSU Fir, Oregon, United States 44.65°N 123.55°W 263 55301 24 0 3.5 - 33.2 0.21 -1.77 6.57
co2_omp_surface-insitu_68_allhours OSU Marys Peak, Oregon, United States 44.50°N 123.55°W 1249 112852 1531 0 1.2 - 18.6 0.64 0.76 3.03
co2_omt_surface-insitu_68_allhours OSU Metolius, Oregon, United States 44.45°N 121.56°W 1255 74274 854 0 1.7 - 16.8 0.57 1.20 3.89
co2_ong_surface-insitu_68_allhours OSU Burns, Oregon, United States 43.47°N 119.69°W 1398 80687 595 0 1.2 - 27.1 0.46 -0.04 3.19
co2_owa_surface-insitu_68_allhours OSU Walton, Oregon, United States 44.07°N 123.63°W 715 37015 847 0 2.0 - 6.6 0.96 -1.03 3.35
co2_oxk_surface-flask_1_representative NOAA Ochsenkopf, Germany 50.03°N 11.81°E 1022 465 7 0 1.4 - 5.7 0.94 -0.97 3.94
co2_oyq_surface-insitu_68_allhours OSU Yaquina Head, Oregon, United States 44.67°N 124.07°W 116 33451 46 0 1.7 - 34.5 0.26 -2.17 4.05
co2_pal_surface-flask_1_representative NOAA Pallas-Sammaltunturi, GAW Station, Finland 67.97°N 24.12°E 565 756 8 0 1.3 - 14.8 0.52 -0.32 2.32
co2_pal_surface-insitu_30_marine FMI Pallas-Sammaltunturi, GAW Station, Finland 67.97°N 24.12°E 565 22767 293 0 1.0 - 3.0 0.81 -0.19 1.14
co2_pal_surface-insitu_30_nonlocal FMI Pallas-Sammaltunturi, GAW Station, Finland 67.97°N 24.12°E 565 94896 2294 0 1.1 - 5.6 0.85 -0.17 1.94
co2_pfa_aircraft-pfp_1_allvalid_0-1000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 551 542 56 0 1.1 - 6.8 1.03 -0.52 3.26
co2_pfa_aircraft-pfp_1_allvalid_1000-2000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 1515 578 46 0 0.4 - 7.1 1.19 -0.34 1.66
co2_pfa_aircraft-pfp_1_allvalid_2000-3000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 2513 647 41 0 0.2 - 4.0 1.33 -0.38 1.22
co2_pfa_aircraft-pfp_1_allvalid_3000-4000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 3467 647 30 0 0.2 - 2.8 1.42 -0.04 1.20
co2_pfa_aircraft-pfp_1_allvalid_4000-5000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 4506 555 51 0 0.3 - 2.2 1.37 0.05 1.05
co2_pfa_aircraft-pfp_1_allvalid_5000-6000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 5446 524 20 0 0.3 - 2.4 1.34 0.06 1.16
co2_pfa_aircraft-pfp_1_allvalid_6000-7000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 6449 511 16 0 0.2 - 2.6 1.37 0.23 1.28
co2_pfa_aircraft-pfp_1_allvalid_7000-8000masl NOAA Poker Flat, Alaska, United States 64.90°N 148.76°W 7209 229 6 0 0.1 - 2.7 1.13 0.18 1.53
co2_poc_shipboard-flask_1_representative NOAA Pacific Ocean variable Surface 2089 74 0 0.0 - 5.3 0.98 -0.06 0.57
co2_prs_surface-insitu_21_allvalid RSE Plateau Rosa Station, Italy 45.93°N 7.70°E 3480 14434 426 0 1.1 - 2.9 0.99 0.14 1.53
co2_psa_surface-flask_1_representative NOAA Palmer Station, Antarctica, United States 64.77°S 64.05°W 10 937 0 0 0.1 - 0.9 0.58 -0.03 0.27
co2_psa_surface-flask_4_representative SIO Palmer Station, Antarctica, United States 64.77°S 64.05°W 10 440 6 0 0.1 - 1.0 0.46 -0.02 0.29
co2_pta_surface-flask_1_representative NOAA Point Arena, California, United States 38.95°N 123.74°W 17 374 2 0 2.8 - 10.8 0.56 -2.34 4.87
co2_px_shipboard-insitu_20_allvalid NIES Pyxis (M/S Pyxis of Toyofuji Shipping Co., Ltd.) variable Surface 317360 24199 0 0.1 - 52.2 1.07 -0.03 1.51
co2_rk1_surface-flask_426_representative SIO_CO2 Kermadec Island, Raoul Island 29.20°S 177.90°W 2 86 6 0 0.2 - 1.2 0.91 -0.25 0.68
co2_rpb_surface-flask_1_representative NOAA Ragged Point, Barbados 13.16°N 59.43°W 15 922 13 0 0.2 - 1.7 1.01 0.06 0.73
co2_rta_aircraft-pfp_1_allvalid_0-1000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 654 103 2 0 0.3 - 0.7 0.98 0.16 0.46
co2_rta_aircraft-pfp_1_allvalid_1000-2000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 1640 309 5 0 0.2 - 0.6 1.13 -0.07 0.47
co2_rta_aircraft-pfp_1_allvalid_2000-3000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 2567 311 3 0 0.3 - 0.8 1.19 -0.15 0.49
co2_rta_aircraft-pfp_1_allvalid_3000-4000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 3483 437 6 0 0.2 - 0.9 1.15 -0.18 0.55
co2_rta_aircraft-pfp_1_allvalid_4000-5000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 4543 328 1 0 0.3 - 0.9 1.23 -0.05 0.58
co2_rta_aircraft-pfp_1_allvalid_5000-6000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 5470 353 1 0 0.2 - 0.9 1.28 -0.12 0.58
co2_rta_aircraft-pfp_1_allvalid_6000-7000masl NOAA Rarotonga, Cook Islands 21.25°S 159.83°W 6246 314 5 0 0.2 - 2.2 0.70 0.04 0.71
co2_san_aircraft-pfp_1_allvalid_1000-2000masl NOAA Santarem, Brazil 2.85°S 54.95°W 1713 10 0 0 1.0 - 4.0 0.65 -1.99 2.44
co2_san_aircraft-pfp_1_allvalid_2000-3000masl NOAA Santarem, Brazil 2.85°S 54.95°W 2527 48 1 0 0.3 - 3.0 0.95 -0.07 1.57
co2_san_aircraft-pfp_1_allvalid_3000-4000masl NOAA Santarem, Brazil 2.85°S 54.95°W 3436 77 0 0 0.5 - 2.2 1.22 0.30 1.02
co2_san_aircraft-pfp_1_allvalid_4000-5000masl NOAA Santarem, Brazil 2.85°S 54.95°W 4600 6 0 0 1.3 - 1.5 0.45 0.42 0.85
co2_sca_aircraft-pfp_1_allvalid_0-1000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 651 285 11 0 0.5 - 4.4 1.16 -0.68 2.61
co2_sca_aircraft-pfp_1_allvalid_1000-2000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 1537 228 6 0 0.6 - 3.5 1.14 0.38 1.77
co2_sca_aircraft-pfp_1_allvalid_2000-3000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 2487 510 20 0 0.3 - 2.7 1.00 0.29 1.30
co2_sca_aircraft-pfp_1_allvalid_3000-4000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 3522 334 13 0 0.2 - 2.8 1.04 0.23 1.17
co2_sca_aircraft-pfp_1_allvalid_4000-5000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 4457 467 13 0 0.2 - 2.3 1.11 0.12 1.01
co2_sca_aircraft-pfp_1_allvalid_5000-6000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 5502 288 7 0 0.4 - 2.0 1.16 0.14 0.88
co2_sca_aircraft-pfp_1_allvalid_6000-7000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 6415 327 5 0 0.4 - 1.5 1.14 0.17 0.83
co2_sca_aircraft-pfp_1_allvalid_7000-8000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 7451 413 12 0 0.3 - 1.5 1.03 0.15 0.77
co2_sca_aircraft-pfp_1_allvalid_8000-9000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 8108 97 2 0 0.2 - 1.8 0.84 0.05 1.21
co2_sca_aircraft-pfp_1_allvalid_9000-10000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 9362 15 0 0 0.3 - 4.8 1.40 -0.34 2.32
co2_sca_aircraft-pfp_1_allvalid_10000-11000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 10432 18 0 0 0.1 - 5.0 1.08 -0.41 1.86
co2_sca_aircraft-pfp_1_allvalid_11000-12000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 11160 7 0 0 0.4 - 5.3 0.85 -0.94 3.25
co2_sca_aircraft-pfp_1_allvalid_12000-13000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 12636 14 0 0 0.3 - 1.6 1.31 0.21 0.88
co2_sca_aircraft-pfp_1_allvalid_13000-14000masl NOAA Offshore Charleston, South Carolina, United States 32.77°N 79.55°W 13145 4 0 0 0.6 - 0.8 1.46 0.17 1.02
co2_sct_surface-pfp_1_allvalid-305magl NOAA Beech Island, South Carolina, United States 33.41°N 81.83°W 115 2556 11 0 2.3 - 33.2 0.40 -0.71 3.54
co2_sct_tower-insitu_1_allvalid-305magl NOAA Beech Island, South Carolina, United States 33.41°N 81.83°W 115 90991 1503 0 2.4 - 23.6 0.83 0.41 4.51
co2_sct_tower-insitu_1_allvalid-31magl NOAA Beech Island, South Carolina, United States 33.41°N 81.83°W 115 15240 449 0 2.7 - 4.8 0.84 -0.60 4.39
co2_sct_tower-insitu_1_allvalid-61magl NOAA Beech Island, South Carolina, United States 33.41°N 81.83°W 115 15592 477 0 2.6 - 4.7 0.84 -0.78 4.11
co2_sey_surface-flask_1_representative NOAA Mahe Island, Seychelles 4.68°S 55.53°E 2 826 0 0 0.3 - 1.0 1.11 -0.11 0.72
co2_sgp_aircraft-pfp_1_allvalid_0-1000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 679 994 42 0 0.2 - 8.3 0.97 -0.16 2.80
co2_sgp_aircraft-pfp_1_allvalid_1000-2000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 1571 1229 44 0 0.6 - 6.2 0.92 0.13 2.09
co2_sgp_aircraft-pfp_1_allvalid_2000-3000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 2462 1197 39 0 0.5 - 2.3 1.01 -0.21 1.43
co2_sgp_aircraft-pfp_1_allvalid_3000-4000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 3467 977 33 0 0.6 - 1.9 1.21 -0.16 1.05
co2_sgp_aircraft-pfp_1_allvalid_4000-5000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 4656 557 22 0 0.4 - 2.1 1.20 -0.18 1.01
co2_sgp_aircraft-pfp_1_allvalid_5000-6000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 5395 236 9 0 0.4 - 1.6 1.29 -0.05 0.90
co2_sgp_aircraft-pfp_1_allvalid_6000-7000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 6464 2 4 0 0.6 - 0.6 3.15 -0.99 0.65
co2_sgp_aircraft-pfp_1_allvalid_8000-9000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 8062 2 0 0 1.3 - 1.3 0.02 -1.20 1.97
co2_sgp_aircraft-pfp_1_allvalid_9000-10000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 9686 7 0 0 0.8 - 0.8 3.31 -0.73 0.63
co2_sgp_aircraft-pfp_1_allvalid_11000-12000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 11321 2 0 0 0.7 - 0.7 0.49 -0.05 0.96
co2_sgp_aircraft-pfp_1_allvalid_12000-13000masl NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 12858 3 0 0 0.5 - 0.5 2.13 -0.18 0.78
co2_sgp_surface-flask_1_representative NOAA Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 314 897 13 0 2.2 - 8.0 0.79 -0.85 3.48
co2_sgp_surface-insitu_64_allvalid-60magl LBNL-ARM Southern Great Plains, Oklahoma, United States 36.61°N 97.49°W 314 19267 664 0 2.4 - 10.0 0.88 -0.40 4.75
co2_shm_surface-flask_1_representative NOAA Shemya Island, Alaska, United States 52.71°N 174.13°E 23 700 15 0 0.7 - 5.9 1.19 -0.50 1.70
co2_sis_surface-flask_2_representative CSIRO Shetland Islands, Scotland 60.09°N 1.25°W 30 82 5 0 0.2 - 2.2 1.36 0.49 1.15
co2_smo_surface-flask_1_representative NOAA Tutuila, American Samoa 14.25°S 170.56°W 42 1663 15 0 0.1 - 2.2 0.78 -0.04 0.39
co2_smo_surface-flask_4_representative SIO Tutuila, American Samoa 14.25°S 170.56°W 42 590 2 0 0.1 - 4.5 0.43 -0.04 0.64
co2_smo_surface-flask_426_representative SIO_CO2 Tutuila, American Samoa 14.25°S 170.56°W 42 665 10 0 0.2 - 2.1 0.76 -0.13 0.52
co2_smo_surface-insitu_1_allvalid NOAA Tutuila, American Samoa 14.25°S 170.56°W 42 36127 0 0 0.2 - 0.8 1.42 -0.01 0.29
co2_spl_surface-insitu_3_nonlocal NCAR Storm Peak Laboratory (Desert Research Institute), United States 40.45°N 106.73°W 3210 63312 1562 0 1.0 - 2.8 1.03 -0.60 1.59
co2_spo_surface-flask_1_representative NOAA South Pole, Antarctica, United States 89.98°S 24.80°W 2810 1430 2 1 0.0 - 1.3 0.23 0.00 0.12
co2_spo_surface-flask_4_representative SIO South Pole, Antarctica, United States 89.98°S 24.80°W 2810 474 2 0 0.1 - 8.9 0.21 -0.01 0.15
co2_spo_surface-flask_426_representative SIO_CO2 South Pole, Antarctica, United States 89.98°S 24.80°W 2810 446 0 0 0.1 - 1.5 0.23 -0.02 0.21
co2_spo_surface-insitu_1_allvalid NOAA South Pole, Antarctica, United States 89.98°S 24.80°W 2810 53887 2 0 0.1 - 0.5 1.56 -0.00 0.09
co2_stm_surface-flask_1_representative NOAA Ocean Station M, Norway 66.00°N 2.00°E 0 780 18 0 0.4 - 5.5 0.99 -0.36 1.26
co2_sum_surface-flask_1_representative NOAA Summit, Greenland 72.60°N 38.42°W 3210 809 59 0 0.3 - 1.2 1.38 -0.12 0.77
co2_svv_tower-insitu_20_allvalid-27magl NIES Savvushka, Russia 51.33°N 82.13°E 495 7513 310 0 1.5 - 5.0 1.13 -0.41 3.76
co2_svv_tower-insitu_20_allvalid-52magl NIES Savvushka, Russia 51.33°N 82.13°E 495 7093 310 0 1.5 - 4.9 1.16 -0.36 3.82
co2_syo_surface-flask_1_representative NOAA Syowa Station, Antarctica, Japan 69.01°S 39.59°E 14 449 0 0 0.1 - 0.6 0.58 -0.09 0.18
co2_tap_surface-flask_1_representative NOAA Tae-ahn Peninsula, Republic of Korea 36.74°N 126.13°E 16 1197 18 0 1.6 - 12.1 0.73 -0.33 5.17
co2_tf1_shipboard-insitu_20_allvalid NIES Trans Future 1 (M/S Trans Future 1 of the Toyofuji Shipping Co., Ltd) variable Surface 75787 1442 0 0.9 - 69.3 0.62 -0.96 3.34
co2_tf5_shipboard-insitu_20_allvalid NIES Trans Future 5 (M/S Trans Future 5 of Toyofuji Shipping Co., Ltd.) variable Surface 399131 10327 0 0.2 - 66.8 0.59 -0.07 1.79
co2_tgc_aircraft-pfp_1_allvalid_0-1000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 692 173 11 0 0.7 - 4.9 0.88 -0.32 2.12
co2_tgc_aircraft-pfp_1_allvalid_1000-2000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 1555 201 4 0 0.5 - 3.1 1.00 0.28 1.43
co2_tgc_aircraft-pfp_1_allvalid_2000-3000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 2521 428 18 0 0.4 - 2.2 0.95 0.14 0.94
co2_tgc_aircraft-pfp_1_allvalid_3000-4000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 3489 250 9 0 0.3 - 1.4 1.10 0.13 0.81
co2_tgc_aircraft-pfp_1_allvalid_4000-5000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 4460 435 11 0 0.2 - 1.3 1.06 0.05 0.76
co2_tgc_aircraft-pfp_1_allvalid_5000-6000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 5552 222 7 0 0.2 - 1.0 1.18 0.10 0.68
co2_tgc_aircraft-pfp_1_allvalid_6000-7000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 6414 246 4 0 0.2 - 1.0 1.22 0.15 0.67
co2_tgc_aircraft-pfp_1_allvalid_7000-8000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 7431 334 6 0 0.1 - 1.1 1.14 0.14 0.63
co2_tgc_aircraft-pfp_1_allvalid_8000-9000masl NOAA Offshore Corpus Christi, Texas, United States 27.73°N 96.86°W 8065 82 1 0 0.3 - 1.3 0.91 0.15 0.73
co2_thd_aircraft-pfp_1_allvalid_0-1000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 629 391 9 0 1.8 - 11.7 0.57 -1.33 5.00
co2_thd_aircraft-pfp_1_allvalid_1000-2000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 1536 213 12 0 0.4 - 2.3 1.34 -0.07 3.83
co2_thd_aircraft-pfp_1_allvalid_2000-3000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 2480 400 15 0 0.2 - 2.4 1.31 0.14 1.30
co2_thd_aircraft-pfp_1_allvalid_3000-4000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 3512 259 19 0 0.1 - 3.8 1.43 0.22 1.43
co2_thd_aircraft-pfp_1_allvalid_4000-5000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 4440 328 13 0 0.2 - 4.4 1.11 0.22 1.34
co2_thd_aircraft-pfp_1_allvalid_5000-6000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 5492 213 6 0 0.2 - 2.5 1.28 0.18 1.12
co2_thd_aircraft-pfp_1_allvalid_6000-7000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 6450 260 12 0 0.3 - 2.2 1.32 0.16 0.99
co2_thd_aircraft-pfp_1_allvalid_7000-8000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 7472 270 17 0 0.2 - 2.0 1.03 0.10 0.98
co2_thd_aircraft-pfp_1_allvalid_8000-9000masl NOAA Trinidad Head, California, United States 41.05°N 124.15°W 8044 22 0 0 0.8 - 1.5 1.13 -0.14 1.20
co2_thd_surface-flask_1_representative NOAA Trinidad Head, California, United States 41.05°N 124.15°W 107 655 3 0 1.9 - 16.3 0.51 -2.29 4.96
co2_tik_surface-flask_1_representative NOAA Hydrometeorological Observatory of Tiksi, Russia 71.60°N 128.89°E 19 280 6 0 1.2 - 6.1 0.86 -0.45 3.72
co2_ulb_aircraft-pfp_1_allvalid_1000-2000masl NOAA Ulaanbaatar, Mongolia 47.40°N 106.00°E 1648 154 4 0 0.3 - 4.3 1.34 -0.01 1.72
co2_ulb_aircraft-pfp_1_allvalid_2000-3000masl NOAA Ulaanbaatar, Mongolia 47.40°N 106.00°E 2471 139 6 0 0.4 - 2.9 1.47 -0.07 1.41
co2_ulb_aircraft-pfp_1_allvalid_3000-4000masl NOAA Ulaanbaatar, Mongolia 47.40°N 106.00°E 3478 147 7 0 0.3 - 2.8 1.57 -0.41 1.50
co2_ulb_aircraft-pfp_1_allvalid_4000-5000masl NOAA Ulaanbaatar, Mongolia 47.40°N 106.00°E 4209 55 2 0 0.1 - 1.8 1.15 -0.47 1.23
co2_ulb_aircraft-pfp_1_allvalid_5000-6000masl NOAA Ulaanbaatar, Mongolia 47.40°N 106.00°E 5718 2 0 0 0.4 - 0.4 2.85 0.55 0.11
co2_ush_surface-flask_1_representative NOAA Ushuaia, Argentina 54.85°S 68.31°W 12 355 3 0 0.2 - 1.0 0.75 -0.22 0.57
co2_uta_surface-flask_1_representative NOAA Wendover, Utah, United States 39.90°N 113.72°W 1327 873 13 1 0.9 - 6.6 0.93 0.75 2.41
co2_uum_surface-flask_1_representative NOAA Ulaan Uul, Mongolia 44.45°N 111.10°E 1007 872 12 0 2.0 - 5.1 0.84 -0.32 3.03
co2_vgn_tower-insitu_20_allvalid-42magl NIES Vaganovo, Russia 54.50°N 62.32°E 192 10293 407 0 2.1 - 4.1 1.13 -0.20 3.40
co2_vgn_tower-insitu_20_allvalid-85magl NIES Vaganovo, Russia 54.50°N 62.32°E 192 10084 375 0 2.0 - 4.9 1.11 -0.25 3.64
co2_wbi_aircraft-pfp_1_allvalid_0-1000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 639 154 8 0 0.8 - 9.8 1.02 -1.11 3.70
co2_wbi_aircraft-pfp_1_allvalid_1000-2000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 1550 382 20 0 0.6 - 6.7 0.87 -0.39 2.70
co2_wbi_aircraft-pfp_1_allvalid_2000-3000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 2550 238 6 0 0.5 - 8.9 1.07 -0.33 1.53
co2_wbi_aircraft-pfp_1_allvalid_3000-4000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 3509 399 17 0 0.3 - 2.2 1.08 -0.10 1.20
co2_wbi_aircraft-pfp_1_allvalid_4000-5000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 4548 283 8 0 0.4 - 2.3 1.31 -0.05 1.11
co2_wbi_aircraft-pfp_1_allvalid_5000-6000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 5520 343 11 0 0.4 - 2.1 1.17 0.01 1.14
co2_wbi_aircraft-pfp_1_allvalid_6000-7000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 6507 308 17 0 0.3 - 1.9 1.40 0.01 1.04
co2_wbi_aircraft-pfp_1_allvalid_7000-8000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 7486 326 10 0 0.3 - 1.7 1.31 0.05 1.07
co2_wbi_aircraft-pfp_1_allvalid_8000-9000masl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 8047 34 2 0 0.5 - 1.2 1.01 -0.14 0.94
co2_wbi_surface-pfp_1_allvalid-379magl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 242 3346 13 0 3.2 - 37.8 0.39 -0.97 4.66
co2_wbi_tower-insitu_1_allvalid-31magl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 242 15667 425 0 3.4 - 9.7 0.85 -0.94 5.00
co2_wbi_tower-insitu_1_allvalid-379magl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 242 99027 2317 0 2.9 - 19.1 0.85 -0.21 4.46
co2_wbi_tower-insitu_1_allvalid-99magl NOAA West Branch, Iowa, United States 41.72°N 91.35°W 242 16570 434 0 3.3 - 9.2 0.82 -1.03 4.83
co2_wgc_tower-insitu_1_allvalid-30magl NOAA Walnut Grove, California, United States 38.27°N 121.49°W 0 16638 280 0 2.7 - 14.0 0.67 -2.54 6.94
co2_wgc_tower-insitu_1_allvalid-483magl NOAA Walnut Grove, California, United States 38.27°N 121.49°W 0 98537 1326 0 2.3 - 12.8 0.64 -0.05 4.58
co2_wgc_tower-insitu_1_allvalid-484magl NOAA Walnut Grove, California, United States 38.26°N 121.49°W 2 1400 16 0 3.3 - 6.7 0.57 -0.31 5.27
co2_wgc_tower-insitu_1_allvalid-89magl NOAA Walnut Grove, California, United States 38.26°N 121.49°W 2 235 1 0 4.6 - 17.2 0.82 -5.92 9.02
co2_wgc_tower-insitu_1_allvalid-91magl NOAA Walnut Grove, California, United States 38.27°N 121.49°W 0 16465 229 0 2.6 - 23.8 0.62 -2.67 6.53
co2_wis_surface-flask_1_representative NOAA Weizmann Institute of Science at the Arava Institute, Ketura, Israel 29.96°N 35.06°E 151 904 9 0 1.3 - 4.2 0.95 -0.57 2.07
co2_wkt_tower-insitu_1_allvalid-122magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 21210 709 0 2.2 - 6.9 0.77 -1.04 3.42
co2_wkt_tower-insitu_1_allvalid-244magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 14848 418 0 2.2 - 6.6 0.97 -0.95 4.00
co2_wkt_tower-insitu_1_allvalid-30magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 21733 675 0 2.3 - 4.2 0.77 -0.85 3.54
co2_wkt_tower-insitu_1_allvalid-457magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 110401 2303 0 2.1 - 40.7 0.77 -0.26 2.96
co2_wkt_tower-insitu_1_allvalid-62magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 2443 86 0 2.0 - 4.7 0.91 -1.14 3.57
co2_wkt_tower-insitu_1_allvalid-9magl NOAA Moody, Texas, United States 31.31°N 97.33°W 251 2851 85 0 2.5 - 6.1 0.89 -0.95 4.34
co2_wlg_surface-flask_1_representative NOAA Mt. Waliguan, Peoples Republic of China 36.29°N 100.90°E 3810 852 9 0 1.0 - 4.3 0.93 -0.11 2.32
co2_wpc_shipboard-flask_1_representative NOAA Western Pacific Cruise variable Surface 182 6 0 0.1 - 2.2 1.07 -0.22 0.66
co2_wsa_surface-insitu_6_allvalid ECCC Sable Island, Nova Scotia, Canada 43.93°N 60.01°W 5 15659 532 0 1.6 - 3.6 1.02 -0.21 2.29
co2_yak_tower-insitu_20_allvalid-11magl NIES Yakutsk, Russia 62.09°N 129.36°E 264 5016 76 0 2.3 - 7.2 0.94 -0.80 5.37
co2_yak_tower-insitu_20_allvalid-77magl NIES Yakutsk, Russia 62.09°N 129.36°E 264 5127 80 0 2.3 - 6.4 0.98 -0.22 5.19
co2_zep_surface-flask_1_representative NOAA Ny-Alesund, Svalbard, Norway and Sweden 78.91°N 11.89°E 474 948 59 0 0.4 - 1.5 1.16 -0.10 0.94
co2_zep_surface-insitu_442_allvalid-15magl ICOS-ATC Ny-Alesund, Svalbard, Norway and Sweden 78.91°N 11.89°E 474 4198 319 0 0.5 - 1.1 1.22 0.00 0.88
co2_zep_surface-insitu_56_allvalid NILU Ny-Alesund, Svalbard, Norway and Sweden 78.91°N 11.89°E 474 1570 217 0 0.4 - 1.7 1.33 0.34 1.05
Summary of Observational Sites Used in CarbonTracker. The site location is specified by latitude, longitude and elevation in meters above sea level. The number of observations actually assimilated for each dataset is listed in the column "Used", and the number rejected due to inability to fit the observations is listed in the column "Rej.". Model-data-mismatch (R) is a value assigned to a given site that is meant to quantify our expected ability to simulate observations there. In this table we report the range of R values assigned to dataset observations by our "adaptive" model-data mismatch scheme (Sec. 7.2). These values are principally determined from the limitations of the atmospheric transport model. It is part of the standard deviation used to interpret the difference between a simulation first guess (Hx) of an observation and the actual measured value (z). The other component, HPHT, is a measure of the ability of the ensemble Kalman filter to improve its simulated value for this observation by adjusting fluxes. These elements together form the innovation χ statistic for the site: χ = (zHx)/√{(HPHT+R2)}. The innovation χ2 reported above is the mean of all squared χ values for a given site. An average χ2 below 1.0 indicates that the HPHT+R2 values are too large. Conversely, values above 1.0 mean that this standard deviation is underestimated. The bias and SE columns are statistics of the posterior residuals (final modeled values - measured values). The bias is the mean of these residuals; the SE is the standard error of those residuals.

8  Ensemble data assimilation

Data assimilation is the process by which a model simulation is adjusted to agree with observations. Model simulations may drift off from reality for a number of reasons. Some models are highly nonlinear, and depend sensitively on knowing the system state with high accuracy. Weather models fall into this category, and as a result reliable forecast systems depend on having a constant stream of meteorological data to correct their simulations. In contrast, models like CarbonTracker need data assimilation not because the controlling dynamics are nonlinear, but because those dynamics are not well known. CarbonTracker uses approximate or estimated rules about the evolution of surface CO2 fluxes, then corrects these approximate projections using observational constraints. The resulting optimal surface flux estimates can then be used to better understand the functioning of the carbon cycle.
Data assimilation is usually a cyclical process, in which estimates get refined over time as more observations become available. Mathematically, data assimilation can be performed using a wide variety of techniques, including variational and ensemble methods. Assimilation systems involving simulations of the global atmosphere are often implemented on highly parallel supercomputers in order to distribute the workload among many computational cores. CarbonTracker is an example of such a model because it relies heavily on estimates of global atmospheric transport.
CarbonTracker model predictions are limited by the relatively simple representations of CO2 surface exchange used to predict land biosphere and ocean fluxes and emissions from fossil fuel combustion and wildfires. As described in the following section, we use data assimilation techniques to modify these surface fluxes so that the resulting atmospheric distribution of CO2 agrees optimally with measurements. We do this by estimating a set of spatially- and temporally-varying scaling factors that multiply first-guess predictions from prior flux models. Data assimilation allows us to determine optimal values for these scaling factors.

8.1  Parameterization of unknowns

CO2 fluxes F(x,y,t) in CarbonTracker are parameterized according to

F(x, y, t) = λ(x, y, t)
Fland(x, y, t) + Focean(x, y, t)
+ FFF(x, y, t) + Ffire(x, y, t),
(11)
where Fland, Focean, FFF, and Fbio are prior flux model predictions for land biosphere, ocean, fossil fuel and wildfire emissions respectively, and λ represents a set of unknown multiplicative 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 prior flux models determine CarbonTracker optimized fluxes. Note that no scaling factors are applied to the fossil fuel and fire modules. The fossil fuel and wildfire fluxes are relatively well-known from prior flux models compared to highly-uncertain land biosphere and ocean fluxes, and as a result we impose those emissions without modification in our model.

8.1.1  Optimization regions

The scaling factors λ are estimated independently for each week and optimization region. They are assumed to be constant over this time period and spatial domain. Each scaling factor is associated with a particular region of the globe, as in the Transcom inversion study (Gurney et al., 2002). Currently the geographic distribution of these optimization regions is fixed. The choice of regions is a strong a priori design decision determining the reliability of the resulting fluxes. In particular, the scale of optimization regions is chosen to minimize "aggregation errors" (2001, 25)Kaminski, Rayner, Heimann, and Enting, while limiting the set of unknown parameters to a manageable number. Following Jacobson et al. (2007), we have divide the global ocean into 30 basins encompassing large-scale ocean circulation and biogeochemical features. The terrestrial biosphere is divided up according to ecosystem type and geographical domain. Specifically, each of the 11 Transcom land regions is subdivided into a maximum of 19 "ecoregions" according to its Olson et al. (1992) vegetation classification. The set of ecoregions over North America is summarized in Table 4 and Figure 18. 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 in Section 9
Theoretically, this approach leads to a total number of 11*19+30=239 optimizable scaling factors for each week, but the actual number of optimization regions is only 156 since some ecosystem types are not represented in every Transcom region. It should be noted also that we have chosen to not optimize scaling factors for ice-covered regions, inland water bodies, and deserts, since the CO2 flux from these regions is negligible.
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° × 1° grid box predominantly covered by coniferous forests will have a different optimized flux λFland(x,y,t) depending on local temperature, radiation, and emissions as simulated by the prior flux model.
Ecosystem types are based on the vegetation classification of Olson et al. (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 categories 23-26+29. Table 4 shows each vegetation category considered. Percentages indicate the relative area in North America associated with each category.
category Olson V 1.3 Percentage area
1 Conifer Forest 19.0%
2 Broadleaf Forest 1.3%
3 Mixed Forest 7.5%
4 Grass/Shrub 12.6%
5 Tropical Forest 0.3%
6 Scrub/Woods 2.1%
7 Semitundra 19.4%
8 Fields/Woods/Savanna 4.9%
9 Northern Taiga 8.1%
10 Forest/Field 6.3%
11 Wetland 1.7%
12 Deserts 0.1%
13 Shrub/Tree/Suc 0.1%
14 Crops 9.7%
15 Conifer Snowy/Coastal 0.4%
16 Wooded tundra 1.7%
17 Mangrove 0.0%
18 Non-optimized areas (ice, polar desert, inland seas) 0.0%
19 Water 4.9%
Table 3: Ecosystem types over North America
Each 1° × 1° pixel of our domain was assigned one of the categories above based on the Olson category that was most prevalent in the 0.5°×0.5° underlying area.

8.1.2  Assimilation window

Measured CO2 mole fractions are the result of upstream surface fluxes and atmospheric transport, which includes both advective movement and diffusive mixing. Near-field surface fluxes can cause significant changes in CO2 mole fractions, whereas flux signals from further upstream become spread out and diluted. Generally speaking, the longer in the past a flux event occurred, the smaller its impact will be on a given sample of air (although it will be spread out through a larger volume of the atmosphere). Thus we choose an "assimilation window" that represents how far back in time we expect to be able to pinpoint a given flux signal from available measurements. A good discussion of this topic can be found in Bruhwiler et al. (2005).
In previous versions of CarbonTracker, the assimilation window was chosen to be five weeks long, meaning that a measurement could cause revisions in surface fluxes only over the 5 weeks leading up to that measurement. In CT2022, we have extended the assimilation window length to 12 weeks. This helps to resolve fluxes in regions of the world with less dense observational coverage (the tropics, Southern Hemisphere, and parts of Asia).
This assimilation window is moved forward on each cycle of our estimation system, so that new weeks are introduced at the "head" of the filter, and the weeks that fall out the "tail" of the filter are finalized. Prior to CT2022, the 5-week assimilation window was moved forward one week at a time. In CT2022, the 12-week assimilation window is moved forward two weeks at a time. Scaling factors λ retain their weekly resolution. Each cycle of the inversion system requires running the atmospheric model for a length of time equal to the assimilation window length plus the window step size. For previous CarbonTracker releases, this was 6 weeks per cycle; for CT2022 it is 14 weeks per cycle. The extra computing time required by the longer assimilation window is balanced somewhat by the two-week stepping, and we have found that CT2022 required only about 15% more computing time per than previous releases.

8.1.3  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). Ensemble statistics are created from 1200 randomly-chosen members, each with its own background CO2 concentration field to represent the time history of that member's surface fluxes. The ensemble Kalman filter looks for correlations between these random flux perturbations and resulting changes in simulated CO2 measurements. We might expect that the entire ensemble would agree that increasing the CO2 flux in a given region results in greater simulated CO2 at a nearby downwind site. However, because we approximate the flux covariance matrix with a random sample of 1200 members, sometimes spurious correlations appear. It is unphysical, for instance, that a measurement at Summit, Greenland could be strongly influenced by surface exchange in the southern Indian Ocean, within the time span of our assimilation window. Any such correlation between the flux ensemble and the measurement in question might be spurious. Localization is a technique developed for numerical weather prediction in which unphysical correlations are diagnosed and systematically ignored (Houtekamer and Mitchell, 1998). We only perform localization for certain datasets. Notably, it is not used for datasets judged to represent hemisphere-scale signals, such as those from marine boundary layer sites in remote locations.
Our localization technique is based on the linear correlation coefficient between the 1200 parameter deviations and 1200 observation deviations for each parameter/observation pair. If the relationship between a parameter deviation and its modeled observational impact is statistically significant, then that relationship is retained. 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.

8.1.4  Dynamical model

In CarbonTracker, the dynamical model is applied to the ensemble-mean parameter values λ as:

λ[t] = (λ0 + λ+[t−1] + λ+[t−2])/3
(12)
Where λ[t] is the prior value of the scaling factors for timestep t, λ0 is the initial prior vector with all elements set to 1.0, and λ+[t−1] and λ+[t−2] are the posterior ("analyzed") scaling factors for timesteps t−1 and t−2 respectively. This model describes that parameter values λ for a new time step are chosen as a combination of optimized values from the two previous time steps and a fixed overall prior value of 1.0. This operation is similar to the simple persistence forecast used in Peters et al. (2005), but represents a smoothing over three time steps, which attenuates variations in the forecast of λ in time. The inclusion of the prior term λ0 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 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 fixed prior covariance structure and not forecast with our dynamical model.

8.2   Covariance structure

The prior covariance structure P0 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 scale 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.
/webdata/ccgg/CT2022/summary/plot_cov.png
Figure 15: CT2022 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 in Sec. 9, 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; 0.5 is equivalent to 50%. The retuning of the covariance matrix for CT2022's multiple-prior simulation (in red) is made evident by also showing these values from previous CarbonTracker releases in light blue and black.
While the correlation structure discussed above has remained fixed in all CarbonTracker releases, we have been changing the overall magnitude of the covariance matrix in an attempt to mitigate seasonal biases in our simulated CO2 fields. Since those biases appear mostly in comparison to measurement data over land, and also because the annual cycle of CO2 in the atmosphere is dominated by the terrestrial carbon cycle, we experimented by loosening the land prior constraint. This is made evident in the "CT2011 through CT2017" trace in the lower panel of Fig. 15. As it turns out, these biases were mostly due to our original short assimilation window length (Sec. 8.1.2) and the convective flux problem discussed in Sec. 6.2. As a result, for CT2022 we were able to scale back the land prior covariance. "L-curve" analysis (Hansen, 1998) suggested that these covariances could be reduced even from the original levels, and new values are shown in red in the lower panel of Fig. 15.

8.3  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 requirement that flux priors be unbiased. We cannot guarantee this for any of our prior 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 CT2022 we present the average of two inversions using different priors. We have used two terrestrial flux priors (including two wildfire emissions estimates), two air-sea CO2 exchange priors, and two estimates of imposed fossil fuel emissions in a factorial design experiment. Whereas previous releases conducted 8 independent inversions exploring each unique combination of priors in a factorial design, we have determined that interactions between priors are negligible. Thus, the quantification of uncertainty due to prior selection can be explored using just two inversions, each using a unique set of priors. We present as a final result the mean flux between these two 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 in its corresponding documentation section.

8.3.1  Posterior uncertainties in CarbonTracker

The formal "internal" error estimates produced by CarbonTracker are unrealistically large. This is largely a result of the dynamical model that introduces a fresh prior covariance matrix with every new week entering the assimilation window. Uncertainties using the new 12-week assimilation window are smaller that those from previous CarbonTracker releases that used a much shorter five-week assimilation window.
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 CT2022 estimates a one-sigma error of about 2.7 Pg C yr−1 on its global flux, this quantity is in actuality much more well-constrained. This is evident from CarbonTracker's excellent agreement both with observational estimates of atmospheric growth rate and with independent CO2 measurements.
In CT2022, our error estimates on optimized fluxes are smaller than those in previous releases. This is mainly due to the retuning of the land prior covariance discussed above. It should be noted that uncertainties presented for CT2022 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.

9  Ecoregions in CarbonTracker

9.1  What are ecoregions?

Ecoregions are the actual scale on which CarbonTracker performs its optimization over land. Ecoregions are meant to represent large expanses of land within a given continent having similar ecosystem types, and are used to divide continent-scale regions into smaller domains for analysis. The ecosystem types use in CarbonTracker are derived from the Olson et al. (1992) vegetation classification (Table 5, Figure 16).
We define an ecoregion as an ecosystem type within a given Transcom land region. There are 19 ecosystem types we extract from the Olson et al. (1992) system, and 11 Transcom land regions (Figure 17), so there are 11 × 19 = 209 possible ecoregions. However, not all ecosystem types are present in all Transcom regions, and the actual number of land ecoregions ends up being 126.
Note on "Semitundra": this is a potentially misleading shorthand abbreviation for a collection of ecosystems comprising semi-desert, shrubs, steppe, and polar+alpine tundra. The "Semitundra" zones appearing in northern Africa where one expects to find the Sahara desert are not, of course, tundra environments. They are instead semi-desert zones.
images/ecotc_glb.png
Figure 16: Global distribution of Olson ecosystem types.
Ecosystem Type North American Boreal North American Temperate
Area (km2) Percentage Area (km2) Percentage
Conifer Forest 2315376 22.9% 1607291 14.0%
Broadleaf Forest - - 269838 2.4%
Mixed Forest 592291 5.9% 930813 8.1%
Grass/Shrub 53082 0.5% 2515582 21.9%
Tropical Forest - - 58401 0.5%
Scrub/Woods - - 416520 3.6%
Semitundra 3396292 33.6% 866468 7.6%
Fields/Woods/Savanna 29243 0.3% 1020939 8.9%
Northern Taiga 1658773 16.4% - -
Forest/Field 61882 0.6% 1243174 10.8%
Wetland 322485 3.2% 66968 0.6%
Deserts - - 21934 0.2%
Shrub/Tree/Suc - - 11339 0.1%
Crops - - 1969912 17.2%
Conifer Snowy/Coastal 41440 0.4% 73437 0.6%
Wooded tundra 360388 3.6% 6643 0.1%
Mangrove - - - -
Non-optimized areas - - - -
Water 1269485 12.6% 384728 3.4%
Total 10100736 100.0% 11463986 100.0%
Table 4: Ecosystem areas over the two Transcom regions covering North America.

9.2  Why use ecoregions?

A fundamental challenge to atmospheric inversions like CarbonTracker is that there are not enough observations to directly constrain fluxes at all times and in all places. It is therefore necessary to find a way to reduce the number of unknowns being estimated. Strategies to reduce the number of unknowns in problems like this one generally impose information from external sources. In CarbonTracker, we reduce the problem size both by estimating fluxes at the ecoregion scale, and by using a terrestrial biological model to give a first guess flux from the ecoregion. The model is also used to give the spatial and temporal distribution of CO2 flux within a region and week.

9.3  Ecosystems within Transcom regions

Each Transcom land region (Figure 17) can contain up to 19 ecoregions.
images/transcom_glb.png
Figure 17: The 11 land regions and 11 ocean regions of the Transcom project, along with the unoptimized land areas in Greenland and Antarctica
.
images/ecotc_nam.png
Figure 18: Ecoregions within the North American Boreal (left) and North American Temperate (right) Transcom regions.
images/ecotc_sam.png
Figure 19: Ecoregions within the South American Tropical (left) and South American Temperate (right) Transcom regions.
images/ecotc_eur.png
Figure 20: Ecoregions within the Europe Transcom region.
images/ecotc_afr.png
Figure 21: Ecoregions within the Northern Africa (left) and Southern Africa (right) Transcom regions.

10  Resources and References

References

[Andres et al. 2011]
R. J. Andres, J. S. Gregg, L. Losey, G. Marland, and T. A. Boden. Monthly, global emissions of carbon dioxide from fossil fuel consumption. Tellus B: Chemical and Physical Meteorology, 63 (3): 309-327, 2011.
[Andres et al. 2014]
R. J. Andres, T. A. Boden, and D. Higdon. A new evaluation of the uncertainty associated with CDIAC estimates of fossil fuel carbon dioxide emission. Tellus B: Chemical and Physical Meteorology, 66 (1): 23616, 2014.
[Baker et al. 2006]
D. F. Baker, R. M. Law, K. R. Gurney, P. Rayner, P. Peylin, A. S. Denning, P. Bousquet, L. Bruhwiler, Y. H. Chen, P. Ciais, I. Y. Fung, M. Heimann, J. John, T. Maki, S. Maksyutov, K. Masarie, M. Prather, B. Pak, S. Taguchi, and Z. Zhu. TransCom 3 CO2 inversion intercomparison: Impact of transport model errors on the interannual variability of regional CO2 fluxes, 1988-2003. Global Biogeochemical Cycles, 20 (1), 2006. GB1002.
[Basu and Nassar 2021]
S. Basu and R. Nassar. Fossil Fuel CO2 Emissions for the OCO2 Model Intercomparison Project (MIP), Jan. 2021. URL https://doi.org/10.5281/zenodo.4776925. Previous versions available from (2018), (2017), (2016), (2015a).
[Basu et al. 2016]
S. Basu, J. B. Miller, and S. Lehman. Separation of biospheric and fossil fuel fluxes of CO2 by atmospheric inversion of CO2 and 14CO2 measurements: Observation system simulations. Atmos. Chem. Phys, 16 (9): 5665-5683, 2016.
[Basu et al. 2020]
S. Basu, S. J. Lehman, J. B. Miller, A. E. Andrews, C. Sweeney, K. R. Gurney, X. Xu, J. Southon, and P. P. Tans. Estimating us fossil fuel CO2 emissions from measurements of 14C in atmospheric CO2. Proceedings of the National Academy of Sciences, 117 (24): 13300-13307, 2020. doi: 10.1073/pnas.1919032117. URL https://www.pnas.org/doi/abs/10.1073/pnas.1919032117.
[Blasing et al. 2004]
T. Blasing, G. Marland, and C. Broniak. Estimates of monthly CO2 emissions and associated 13C/12C values from fossil-fuel consumption in the USA (1981-2003). Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., U.S.A., 2004. doi: 10.3334/CDIAC/ffe.001.
[Boden et al. 2017]
T. A. Boden, G. Marland, and R. J. Andres. Global, regional, and national fossil fuel CO2 emissions. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., U.S.A., 2017. doi: 10.3334/CDIAC/00001_V2017.
[BP 2021]
BP. BP Statistical Review of World Energy. BP p.l.c., London, 2021. URL https://www.bp.com/en/global/corporate/energy-economics/statistical-review-of-world-energy.html.
[British Petroleum 2019]
British Petroleum. BP Statistical Review of World Energy. Number 68. BP p.l.c., London, 2019.
[Bruhwiler et al. 2005]
L. M. P. Bruhwiler, A. M. Michalak, W. Peters, D. F. Baker, and P. Tans. An improved Kalman smoother for atmospheric inversions. Atmospheric Chemistry and Physics, 5: 2691-2702, 2005. doi: https://doi.org/10.5194/acp-5-2691-2005.
[Byrne et al. 2022]
B. Byrne, D. F. Baker, S. Basu, M. Bertolacci, K. W. Bowman, D. Carroll, A. Chatterjee, F. Chevallier, P. Ciais, N. Cressie, D. Crisp, S. Crowell, F. Deng, Z. Deng, N. M. Deutscher, M. Dubey, S. Feng, O. García, D. W. T. Griffith, B. Herkommer, L. Hu, A. R. Jacobson, R. Janardanan, S. Jeong, M. S. Johnson, D. B. A. Jones, R. Kivi, J. Liu, Z. Liu, S. Maksyutov, J. B. Miller, S. M. Miller, I. Morino, J. Notholt, T. Oda, C. W. O'Dell, Y.-S. Oh, H. Ohyama, P. K. Patra, H. Peiro, C. Petri, S. Philip, D. F. Pollard, B. Poulter, M. Remaud, A. Schuh, M. K. Sha, K. Shiomi, K. Strong, C. Sweeney, Y. Té, V. A. Velazco, M. Vrekoussis, T. Warneke, et al. National CO2 budgets (2015-2020) inferred from atmospheric CO2 observations in support of the Global Stocktake. Earth System Science Data Discussions, 2022: 1-59, 2022. doi: 10.5194/essd-2022-213. URL https://essd.copernicus.org/preprints/essd-2022-213/.
[Caldeira and Wickett 2003]
K. Caldeira and M. E. Wickett. Anthropogenic carbon and ocean pH. Nature, 425 (6956): 365-365, 2003.
[Commission et al. 2019]
E. Commission, J. R. Centre, F. Monforti-Ferrario, G. Oreggioni, E. Schaaf, D. Guizzardi, J. Olivier, E. Solazzo, E. Lo Vullo, M. Crippa, M. Muntean, and E. Vignati. Fossil CO2 and GHG emissions of all world countries : 2019 report. Publications Office, 2019. doi: doi/10.2760/687800. URL https://data.jrc.ec.europa.eu/collection/EDGAR.
[Giglio et al. 2006]
L. Giglio, G. R. van der Werf, J. T. Randerson, G. J. Collatz, and P. Kasibhatla. Global estimation of burned area using MODIS active fire observations. Atmospheric Chemistry and Physics, 6: 957-974, 2006.
[Giglio et al. 2013]
L. Giglio, J. T. Randerson, and G. R. van der Werf. Analysis of daily, monthly, and annual burned area using the fourth-generation global fire emissions database (GFED4). Journal of Geophysical Research: Biogeosciences, 118 (1): 317-328, 2013. doi: 10.1002/jgrg.20042. URL https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1002/jgrg.20042.
[Gilfillan and Marland 2021]
D. Gilfillan and G. Marland. CDIAC-FF: global and national CO2 emissions from fossil fuel combustion and cement manufacture: 1751-2017. Earth System Science Data, 13 (4): 1667-1680, 2021. doi: 10.5194/essd-13-1667-2021. URL https://essd.copernicus.org/articles/13/1667/2021/.
[Gloor et al. 2003]
M. Gloor, N. Gruber, J. Sarmiento, C. L. Sabine, R. A. Feely, and C. Rödenbeck. A first estimate of present and preindustrial air-sea CO2 flux patterns based on ocean interior carbon measurements and models. Geophysical Research Letters, 30 (1): 10.1029/2002GL015594, 2003.
[Gruber et al. 1996]
N. Gruber, J. L. Sarmiento, and T. F. Stocker. An improved method for detecting anthropogenic CO2 in the oceans. Global Biogeochemical Cycles, 10 (4): 809-837, 1996.
[Gurney et al. 2002]
K. R. Gurney, R. M. Law, A. S. Denning, P. J. Rayner, D. Baker, P. Bousquet, L. Bruhwiler, Y.-H. Chen, P. Ciais, S. Fan, I. Y. Fung, M. Gloor, M. Heimann, K. Higuchi, J. John, T. Maki, S. Maksyutov, K. Masarie, P. Peylin, M. Prather, B. Pak, J. Randerson, J. L. Sarmiento, S. Taguchi, T. Takahashi, P. Tans, and C.-W. Yuen. Towards robust regional estimates of CO2 sources and sinks using atmospheric transport models. Nature, 415, 2002.
[Hansen 1998]
P. C. Hansen. Rank-deficient and discrete ill-posed problems: numerical aspects of linear inversion. SIAM, 1998.
[Houtekamer and Mitchell 1998]
P. L. Houtekamer and H. L. Mitchell. Data assimilation using an ensemble Kalman filter technique. Monthly Weather Review, 126 (3): 796-811, 1998. doi: 10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2. URL https://doi.org/10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2.
[Jacobson et al. 2007]
A. R. Jacobson, N. Gruber, J. L. Sarmiento, M. Gloor, and S. E. M. Fletcher. A joint atmosphere-ocean inversion for surface fluxes of carbon dioxide: I. methods and global-scale fluxes. Global Biogeochemical Cycles, 21 (GB1019), 2007.
[Jacobson et al. 2023]
A. R. Jacobson, K. N. Schuldt, P. Tans, A. Andrews, J. B. Miller, T. Oda, S. Basu, J. Mund, B. Weir, L. Ott, T. Aalto, J. B. Abshire, K. Aikin, S. Aoki, F. Apadula, S. Arnold, B. Baier, J. Bartyzel, A. Beyersdorf, T. Biermann, S. C. Biraud, H. Boenisch, G. Brailsford, W. A. Brand, G. Chen, H. Chen, L. Chmura, S. Clark, A. Colomb, et al. CarbonTracker CT20222. Model published by NOAA Global Monitoring Laboratory, 2023. URL https://gml.noaa.gov/ccgg/carbontracker/CT2022/.
[Kaminski et al. 2001]
T. Kaminski, P. J. Rayner, M. Heimann, and I. G. Enting. On aggregation errors in atmospheric transport inversions. Journal of Geophysical Research-Atmospheres, 106 (D5): 4703-4715, 2001.
[Krol et al. 2005]
M. Krol, S. Houweling, B. Bregman, M. van den Broek, A. Segers, P. van Velthoven, W. Peters, F. Dentener, and P. Bergamaschi. The two-way nested global chemistry-transport zoom model TM5: algorithm and applications. Atmospheric Chemistry and Physics, 5: 417-432, 2005. URL http://www.atmos-chem-phys.net/5/417/2005/acp-5-417-2005.html.
[Levitus et al. 2010]
S. Levitus, R. A. Locarnini, T. P. Boyer, A. V. Mishonov, J. I. Antonov, H. E. Garcia, O. K. Baranova, M. M. Zweng, D. R. Johnson, and D. Seidov. World Ocean Atlas 2009. 2010.
[Marland 2008]
G. Marland. Uncertainties in accounting for CO2 from fossil fuels. Journal of Industrial Ecology, 12 (2): 136-139, 2008. doi: 10.1111/j.1530-9290.2008.00014.x. URL https://onlinelibrary.wiley.com/doi/abs/10.1111/j.1530-9290.2008.00014.x.
[Mu et al. 2011]
M. Mu, J. Randerson, G. Van der Werf, L. Giglio, P. Kasibhatla, D. Morton, G. Collatz, R. DeFries, E. Hyer, E. Prins, et al. Daily and 3-hourly variability in global fire emissions and consequences for atmospheric model predictions of carbon monoxide. Journal of Geophysical Research: Atmospheres, 116 (D24), 2011.
[Nassar et al. 2013]
R. Nassar, L. Napier-Linton, K. R. Gurney, R. J. Andres, T. Oda, F. R. Vogel, and F. Deng. Improving the temporal and spatial distribution of CO2 emissions from global fossil fuel emission data sets. Journal of Geophysical Research: Atmospheres, 118 (2): 917-933, 2013. doi: 10.1029/2012JD018196. URL https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2012JD018196.
[Oda and Maksyutov 2011]
T. Oda and S. Maksyutov. A very high-resolution (1 km x 1 km) global fossil fuel CO2 emission inventory derived using a point source database and satellite observations of nighttime lights. Atmospheric Chemistry and Physics, 11 (2): 543-556, 2011. doi: 10.5194/acp-11-543-2011. URL https://www.atmos-chem-phys.net/11/543/2011/.
[Oda and Maksyutov 2015]
T. Oda and S. Maksyutov. ODIAC fossil fuel CO2 emissions dataset. doi:10.17595/20170411.001, 2015.
[Oda et al. 2018]
T. Oda, S. Maksyutov, and R. J. Andres. The open-source data inventory for anthropogenic CO2, version 2016 (ODIAC2016): a global monthly fossil fuel CO2 gridded emissions data product for tracer transport simulations and surface flux inversions. Earth System Science Data, 10 (1): 87-107, 2018. doi: 10.5194/essd-10-87-2018. URL https://www.earth-syst-sci-data.net/10/87/2018/.
[Olsen and Randerson 2004]
S. C. Olsen and J. T. Randerson. Differences between surface and column atmospheric CO2 and implications for carbon cycle research. Journal of Geophysical Research-Atmospheres, 109 (D2), 2004.
[Olson et al. 1992]
J. Olson, J. Watts, and L. Allsion. Major world ecosystem complexes ranked by carbon in live vegetation: A database. Technical Report ORNL/CDIAC-134, NDP-017, 1992. URL http://cdiac.ornl.gov/epubs/ndp/ndp017/carbonbig.html.
[Pacanowski and Gnanadesikan 1998]
R. C. Pacanowski and A. Gnanadesikan. Transient response in a z-level ocean model that resolves topography with partial cells. Monthly Weather Review, 126: 3248-3270, 1998.
[Patra et al. 2011]
P. K. Patra, S. Houweling, M. Krol, P. Bousquet, D. Belikov, D. Bergmann, H. Bian, P. Cameron-Smith, M. P. Chipperfield, K. Corbin, et al. TransCom model simulations of CH4 and related species: linking transport, surface flux and chemical loss with CH4 variability in the troposphere and lower stratosphere. Atmospheric Chemistry and Physics, 11 (24): 12813-12837, 2011.
[Peters et al. 2004]
W. Peters, M. C. Krol, E. J. Dlugokencky, F. J. Dentener, P. Bergamaschi, G. Dutton, P. von Velthoven, J. B. Miller, L. Bruhwiler, and P. P. Tans. Toward regional-scale modeling using the two-way nested global model TM5: Characterization of transport using SF6. Journal of Geophysical Research-Atmospheres, 109 (D19), 2004. D19314.
[Peters et al. 2005]
W. Peters, J. Miller, J. Whitaker, A. Denning, A. Hirsch, M. Krol, D. Zupanski, L. Bruhwiler, and P. Tans. An ensemble data assimilation system to estimate CO2 surface fluxes from atmospheric trace gas observations. Journal of Geophysical Research-Atmospheres, 110: D24304, Jan 2005. doi: 10.1029/2005JD006157. D24304.
[Potter et al. 1993]
C. S. Potter, J. T. Randerson, C. B. Field, P. A. Matson, P. M. Vitousek, H. A. Mooney, and S. A. Klooster. Terrestrial ecosystem production - a process model-based on global satellite and surface data. Global Biogeochemical Cycles, 7 (4): 811-841, 1993.
[Rasmussen 1991]
L. Rasmussen. Piecewise integral splines of low degree. Computers & Geosciences, 17 (9): 1255-1263, 1991.
[Sabine et al. 2004]
C. L. Sabine, R. A. Feely, N. Gruber, R. M. Key, K. Lee, J. L. Bullister, R. Wanninkhof, C. S. Wong, D. W. R. Wallace, B. Tilbrook, F. J. Millero, T. H. Peng, A. Kozyr, T. Ono, and A. F. Rios. The oceanic sink for anthropogenic CO2. Science, 305 (5682): 367-371, 2004.
[Schuh et al. 2019]
A. E. Schuh, A. R. Jacobson, S. Basu, B. Weir, D. Baker, K. Bowman, F. Chevallier, S. Crowell, K. J. Davis, F. Deng, et al. Quantifying the impact of atmospheric transport uncertainty on CO2 surface flux estimates. Global Biogeochemical Cycles, 33 (4): 484-500, 2019.
[Takahashi et al. 2002]
T. Takahashi, S. C. Sutherland, C. Sweeney, A. P. N. Metzl, B. Tilbrook, N. Bates, R. Wanninkhof, R. A. Feely, C. Sabine, J. Olafsson, and Y. Nojiri. Global air-sea CO2 flux based on climatological surface ocean pCO2, and seasonal biological and temperature effects. Deep-Sea Research II, 49: 1601-1622, 2002.
[Takahashi et al. 2009]
T. Takahashi, S. C. Sutherland, R. Wanninkhof, C. Sweeney, R. A. Feely, D. W. Chipman, B. Hales, G. Friederich, F. Chavez, C. Sabine, et al. Climatological mean and decadal change in surface ocean pCO2, and net sea-air CO2 flux over the global oceans. Deep Sea Research Part II: Topical Studies in Oceanography, 56 (8-10): 554-577, 2009.
[Thoning et al. 1989]
K. W. Thoning, P. P. Tans, and W. D. Komhyr. Atmospheric carbon dioxide at Mauna Loa observatory. 2. Analysis of the NOAA GMCC data, 1974-1985. Journal of Geophysical Research-Atmospheres, 94: 8549-8565, Jan 1989.
[van der Werf et al. 2003]
G. van der Werf, J. Randerson, G. Collatz, and L. Giglio. Carbon emissions from fires in tropical and subtropical ecosystems. Global Change Biology, 9: 547-562, Jan 2003.
[van der Werf et al. 2006]
G. R. van der Werf, J. T. Randerson, L. Giglio, G. J. Collatz, P. S. Kasibhatla, and A. F. Arellano. Interannual variability in global biomass burning emissions from 1997 to 2004. Atmospheric Chemistry and Physics, 6: 3423-3441, Jan 2006.
[van der Werf et al. 2017]
G. R. van der Werf, J. T. Randerson, L. Giglio, T. T. van Leeuwen, Y. Chen, B. M. Rogers, M. Mu, M. J. E. van Marle, D. C. Morton, G. J. Collatz, R. J. Yokelson, and P. S. Kasibhatla. Global fire emissions estimates during 1997-2016. Earth System Science Data, 9 (2): 697-720, 2017. doi: 10.5194/essd-9-697-2017. URL https://www.earth-syst-sci-data.net/9/697/2017/.
[Wanninkhof 1992]
R. Wanninkhof. Relationship between wind-speed and gas-exchange over the ocean. Journal of Geophysical Research-Oceans, 97: 7373-7382, Jan 1992.
[Whitaker and Hamill 2002]
J. S. Whitaker and T. M. Hamill. Ensemble data assimilation without perturbed observations. Monthly Weather Review, 130 (7): 1913-1924, 2002.