Climate Change 2021: The Physical Science Basis
Climate Change 2021: The Physical Science Basis
Edited by
ISBN 978-92-9169-163-0
Front cover artwork: Changing by Alisa Singer, www.environmentalgraphiti.org © 2021 Alisa Singer.
Foreword, Preface
and Dedication
Foreword
Foreword
Foreword
It is unequivocal that human activities have heated our climate. The journey of this report reflects extraordinary efforts by all
Recent changes are rapid, intensifying, and unprecedented over contributors under exceptional circumstances. The COVID-19
centuries to thousands of years. With each additional increment of pandemic, which started during the review of the Second Order
warming, these changes will become larger, resulting in long-lasting, Draft, disrupted the process. However, the early response of Working
irreversible implications, in particular for sea level rise. United Nations Group I, through the hard work of the authors, review editors, chapter
Secretary-General António Guterres has stated that ‘the evidence is scientists, the Technical Support Unit, and Bureau members, made
irrefutable’ and ‘we see the warning signs in every continent and region’. it possible to deliver a report that meets the most stringent science
quality standards, thanks to the in-depth, broad review process, with
The Working Group I contribution to the Intergovernmental Panel an exhaustive, robust, rigorous and transparent assessment of the
on Climate Change (IPCC) Sixth Assessment Report provides a reality latest state of climate science knowledge. We commend the additional
check on climate change. We now have a much clearer picture of efforts taken, including by all reviewers from the research community
the past, present and possible future climates, and this information and government representatives who adapted to a new way of working
is essential for understanding where we are headed, what can be to complete this vital report. Our particular appreciation goes to the
done, and the multiple facets of a changing climate to prepare for, Working Group I Co-Chairs Valérie Masson-Delmotte and Panmao Zhai
in every region. Unless deep reductions in greenhouse gas emissions for their leadership throughout the process. Their extraordinary efforts
occur in the coming decades, global warming of 1.5°C and 2°C above ensured that this robust and clear report is available to the world today.
pre-industrial levels will be exceeded during the 21st century.
This Working Group I contribution to the IPCC’s Sixth Assessment
The new Working Group I Report structure integrates multiple lines Report covers important new advances in climate science that
of scientific evidence within each chapter and provides robust provide an invaluable input into climate negotiations and decision-
knowledge relevant for policymaking through the end-to-end making, with an emphasis on key conditions that, from a physical
assessment of key topics. This more holistic approach has resulted in science basis, are needed to limit global warming and inform risk
a report that provides a better understanding of the climate system, assessment and regional adaptation. The report was welcomed at
for both past and future changes, with a new emphasis on climate the 26th Conference of the Parties to the United Nations Framework
information for regions, which is critical for informing adaptation and Convention on Climate Change. It will also inform the 2023 global
risk management strategies. Today, many decisions remain grounded stocktake. Unless there are deep reductions in global greenhouse gas
in the experience of past climate variability. This report provides emissions, the goal of limiting warming well below 2°C and close to
a solid basis for taking into account future changes that need to be 1.5°C will be out of reach.
considered in today’s decisions, with increased relevance for climate
services, to enhance adaptation and resilience to climate change and The science is unequivocal, the changes are unprecedented, and there
curb greenhouse gas emissions. is no more time for delay.
v
Preface
Preface
Preface
The Working Group I (WGI) contribution to the Sixth Assessment without being policy prescriptive and to facilitate the integration
Report (AR6) of the Intergovernmental Panel on Climate Change of the WGI key findings with the other AR6 Working Group reports.
(IPCC) focuses on a full and comprehensive assessment of the physical
science basis of climate change, based on evidence from more than
14,000 scientific publications available by 31 January 2021. Structure of the Report
This Report reflects recent climate science advances resulting from This Report consists of thirteen thematic chapters with their
progress in, and the integration of, multiple lines of evidence, supporting supplementary material, ten annexes (including the report
including: in situ and remote observations; paleoclimate information; Glossary, which is developed in coordination with Working Groups
understanding of climate drivers and physical, chemical and biological II and III where relevant), an integrative Technical Summary, and
processes and feedbacks; and global and regional climate modelling; a Summary for Policymakers. An innovation in this Working Group I
as well as advances in methods of analyses and insights from the assessment is the online Interactive Atlas (https://interactive-atlas.
growing field of climate services. ipcc.ch), a novel tool for flexible spatial and temporal analyses of
observed and projected climate change information, which enhances
The AR6 WGI Report builds on the WGI contribution to the IPCC’s accessibility for stakeholders and users to the data assessed in
Fifth Assessment Report (AR5) in 2013, and the AR6 Special Reports1 the Report.
released in 2018 and 2019.
The Summary for Policymakers and Technical Summary include the line
The report considers the current state of the climate in the long-term of sight, indicated in curly brackets, to the chapters and the specific
context, the understanding of human influence, the state of knowledge sections therein where the detailed assessment can be found. In this
about possible climate futures, climate information relevant for way, these summary components of the Report provide a roadmap to
climate-related risk assessment and regional adaptation, and the the contents of the entire report and a synthesis of the major findings
physical science basis on limiting human-induced climate change. that is traceable to the underlying literature and assessment.
1 Global warming of 1.5°C: an IPCC special report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context
of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty (SR1.5); Climate Change and Land: an IPCC special report on
climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems (SRCCL); IPCC Special Report on the Ocean and
Cryosphere in a Changing Climate (SROCC).
vii
Preface
oxide and the assessment of carbon and other biogeochemical technical aspects of the assessment, such as descriptions of datasets,
feedbacks. Chapter 6 assesses changes in the emissions and models or methodologies supporting chapter analyses.
Preface
viii
Preface
science journals, the timeline of the report was extended by four The Summary for Policymakers was approved line-by-line during the
months to balance the delays and challenges faced by authors first-ever virtual IPCC approval session, the 14th Session of IPCC
Preface
and the scientific community with maintaining momentum and Working Group I from 26 July – 06 August 2021, and the underlying
timeliness in the assessment process. The review process is critical Report was accepted during the 54th Session of the IPCC on the
for the rigor, objectivity and comprehensiveness of the assessment. 6th August 2021.
The adjustments to the timeline facilitated broad participation from
scientists and governments in the review process. Stringent scientific
rigor and quality of the assessment were maintained despite Acknowledgements
the pandemic.
We are very grateful for the exceptional rigor and dedication of the
The fourth Lead Author Meeting, due to be held in June 2020, volunteer Coordinating Lead Authors and Lead Authors throughout
was replaced by extensive virtual activities to address the Second this process, who delivered the most comprehensive assessment
Order Draft review comments and topics that cut across multiple ever of our physical understanding of climate change. We thank the
chapters. A final virtual Lead Author Meeting was held in February Review Editors for working alongside the author teams to ensure
2021 to finalize the report. Drafting meetings for the Summary for that the chapters are fully reflective of the input provided through
Policymakers also took place through virtual meetings. the review process. We express our sincere appreciation to all the
government and expert reviewers, including several group reviews
Addressing the implications of the COVID-19 pandemic on the from early-career scientists. We thank the many Contributing Authors
assessment process required innovation to facilitate international who provided input and important support to the authors.
virtual collaboration, including extra support and training for
participants and facilitators, support for participants with internet A special thanks goes to the Chapter Scientists of this report who
connectivity challenges, additional advance preparation, shorter went above and beyond what was expected of them: Kari Alterskjaer,
and more focused meetings with clear agendas and objectives and Lisa Bock, Katherine Dooley, Gregory Garner, Mathias Hauser, Tim
duplicated to account for time zones of participants, high levels Hermans, Lijuan Hua, Carley Iles, Maialen Iturbide, Laurice Preciado
of transparency – including the provision of written summaries Jamero, Martin Jury, Megan Kirchmeier Young, Chaincy Kuo, Hui-Wen
of meetings and decisions, and allotting time for asynchronous Lai, Alice Lebehot, Elizaveta Malinina Rieger, Sebastian Milinkski,
contributions to the discussion and decision-making process. Therese Myslinski, Tamzin Palmer, Browdie Pearson, Stephane Senesi,
Jérôme Servonnat, Chris Smith, David Smyth, Sabin Thazhe Purayil,
Documenting and understanding barriers to participation due to Emilie Vanvyve, Tania Villasenor Jorquera, Hui Wan and Kyung-
an increased reliance on online activities and the use of inclusive Sook Yu. Chapter scientists were recruited by and reported directly
practices required priority attention in these novel conditions. We to the Coordinating Lead Author(s) and provided technical support
gained experience in applying methods to facilitate participatory and to the chapters, including reference checking and compilation,
inclusive processes in the assessment and recognized the necessity figure drafting, traceability checking, identification of overlaps or
of fostering these approaches over the course of the assessment inconsistencies across chapters, and technical editing.
process, both during and in-between meetings, to build a stronger
community of practice within this unique international context. This We thank the Vice Chairs of the WGI Bureau for their dedication,
will be an important legacy for future assessment cycles. guidance and wisdom throughout the preparation of the Report and
their support for cross-Working Group coordination: Edvin Aldrian,
The preparation of the WGI AR6 report was also informed by Fatima Driouech, Gregory Flato, Jan Fuglestvedt, Muhammad I. Tariq,
recommendations from several IPCC expert meetings. The first Carolina Vera and Noureddine Yassaa.
meeting focused on assessing climate information for regions
(in 2018, co-organized by WGI and WGII, and hosted at ICTP, Trieste, We gratefully acknowledge the support from the host countries
Italy), which provided a scoping of the Interactive Atlas. A second and institutions of the WGI Lead Author Meetings (LAMs): China
meeting focused on short-lived climate forcers (in 2018, co-organized Meteorological Administration (CMA), China, for hosting the first
by TFI and WGI, and hosted by the World Meteorological Organization LAM; Environment Canada, Canada, for hosting the second LAM; and
in Geneva, Switzerland), which identified science advances in the Météo France, France, for hosting the third LAM. We also thank the
understanding of emissions and climate effects of SLCFs and needs Ministerio de Ciencia, Tecnologia, Conocimiento e Innovación and
for improvements in emission inventories and methodologies. A the Ministerio del Medio Ambiente, Chile, for offering to host the
third meeting on mitigation, sustainability, and climate stabilization fourth LAM meeting that we were unable to hold in person due to
scenarios (in 2019, organized by WGIII in Addis Ababa, Ethiopia), led the COVID-19 pandemic.
to cross-WG coordination related to scenarios. Recommendations
for clarity and readability from the 2016 IPCC Expert Meeting The support provided by governments and institutions, as well
on Communication (organized by the IPCC Secretariat in Oslo, as through contributions to the IPCC Trust Fund, is thankfully
Norway) were taken into account in developing technical guidance, acknowledged as it enabled the participation of the author
training and resources provided to authors and in particular for the teams in the preparation of the Report. The efficient operation of
preparation of the text and figures of the Frequently Asked Questions the WGI Technical Support Unit (TSU) was made possible by the
and the Summary for Policymakers. generous financial support provided by the government of France
ix
Preface
and administrative and information technology support from the Lynn, Judith Ewa, Mxolisi Shongwe, Jennifer Lew Schneider, Jesbin
Université Paris Saclay (France), Institut Pierre Simon Laplace (IPSL) Baidya, Werani Zabula, Nina Peeva, Melissa Walsh, Joelle Fernandez,
Preface
and the Laboratoire des Sciences du Climat et de l’Environnement Laura Biagioni and Oksana Ekzarkho for their guidance and support
(LSCE). We thank the Norwegian Environment Agency for supporting to implement the many facets of the IPCC process. We were grateful
the preparation of the graphics for the Summary for Policymakers. that we could work together extensively on communication activities.
We thank the United Nations Environment Programme Library for We thank Sue Escott for her tireless work to strengthen how we
providing a service for authors to access literature. communicate the outcomes of the assessment.
The approval of the WGI Summary for Policymakers took place in We would like to acknowledge the support of the SHIFT Collaborative
an unprecedented context, with travel restrictions caused by the team – Stacy Barter and Michelle Colussi – and the generous
COVID-19 pandemic rendering an in-person IPCC Plenary session support of the Canadian government for training and tools on
impossible. We thank the support and advice of the IPCC Executive inclusive practices in a consensus-based decision-making context,
Committee and the tireless work of the ad-hoc task group that was which we have been able to use for a more inclusive assessment
established to advise us as Co-Chairs in preparing for the approval process, including when we moved to purely online approaches. We
session. The task group was led by IPCC Vice-Chair Ko Barrett and appreciated the presence of Jessica O’Reilly and Mark Vardy, who
included WGI Vice-Chairs Fatima Driouech, Greg Flato and Edvin have been with us throughout, working on an ethnographic study of
Aldrian; Anna Pirani and Sarah Connors of the WGI TSU; and Ermira how authors undertake the IPCC assessment, and we look forward to
Fida of the IPCC Secretariat. The task group prepared guidance for the insights from their research.
participants on the modalities of the session and a carefully structured
meeting schedule to implement a virtual approval process. A core outcome of the report has been the development of the WGI
Interactive Atlas produced by the Atlas chapter team. The Atlas is
The approval took place virtually for the first time and involved more dedicated to the memory of Gemma Teresa Narisma, who co-led this
than 186 hours of online meetings. We thank all participants for the innovative chapter with her extensive experience in regional climate
remarkable collaborative spirit and rigorous work undertaken during research and outstanding leadership. The development and technical
the session. IPCC Vice Chairs Ko Barrett, Thelma Krug and Youba implementation have been supported with in-kind contribution
Sokona brought their unwavering support to facilitate discussions from the Spanish government through the Spanish Research
amongst authors and delegations and provided core support for the Council (CSIC) Instituto de Física de Cantabria, in partnership with
success of the approval process. We are also grateful to the Vice Chairs Predictia Intelligent Data Solutions. Funding from the Spanish
of WGI, as well as Mark Howden and Andy Reisinger, Vice Chairs of Research & Development program is acknowledged (ref. PID2019-
WGII and WGIII respectively, and Jim Skea, Co-Chair of WGIII, for 111481RB-I00). We thank the modelling centres and institutions
their support to facilitate the discussions. The WGI TSU was joined that produce and make available the datasets used in this work. The
by members of the WGII and WGIII TSUs, as well as past interns and Interactive Atlas was first made available along with the rest of the
chapter scientists to staff this Herculean coordination effort. Report on 9 August 2021 and was visited by more than half a million
users worldwide during the first month.
Our warmest thanks go to the collegial and collaborative support
provided by Melinda Tignor, Elvira Poloczanska, Katja Mintenbeck, The WGI TSU has initiated the process to archive the data and
Bard Rama, Almut Niebuhr, Vincent Möller, Sina Löschke, Komila code from the report, building on the guidance and support from
Nabiyeva, Andrés Alegría, Stefanie Langsdorf, Andrew Okem, a large group of contributors. We are indebted to the members
Marlies Craig, Anka Mühle, Philisiwe Manqele, Stefan Weisfeld, of the IPCC Task Group on Data Support for Climate Change
Jussi Savolainen and Mallou from the Working Group II Technical Assessments (TG-Data) for their oversight, expert guidance and
Support Unit; Roger Fradera, Raphael Slade, Alaa Al Khourdajie, Minal constant encouragement, including the Co-chairs of the Task Group,
Pathak, Sigourney Luz, Malek Belkacemi, David McCollum, Renée van David Huard and Sebastian Vicuna, and members representing the
Diemen, Shreya Some, Purvi Vyas, Juliette Malley and Géninha Lisboa WGI science community, Michio Kawamiya, Silvina Solman, José
from the Working Group III Technical Support Unit; and Noémie Manuel Guttierez and Nana Ama Browne Klutse. For the preparation
Le Prince-Ringuet from the Synthesis Report Technical Support Unit. of the figure data and code for archival, we especially thank to our
dedicated contractor, Lina Sitz.
We are grateful for the close collaboration with authors and Bureau
members from Working Group II and III, including as contributing The IPCC Data Distribution Centre (DDC) has been indispensable
authors in many parts of the report. We thank the Co-Chairs Debra for this effort. For the archival of figure data, we are indebted to
Roberts, Hans-Otto Portner, Jim Skea and Priyadarshi R. Shukla for Charlotte Pascoe, Kate Winfield, and Martin Jukes from the UK
the collegial teamwork across Working Groups that has characterized Centre for Environmental Data Analysis (CEDA). For the archival of
the AR6. We also thank Eduardo Calvo Buendía and Kyoto Tanabe, the climate model data used as input to the report and intermediate
Co-Chairs of the Task Force on Greenhouse Gas Inventories, for their assessed datasets, we gratefully acknowledge Martina Stockhause
support and collaboration. of the German Climate Computing Centre (DKRZ). For the transfer
of metadata on archived data/code into the IPCC data catalogue, we
We thank Hoesung Lee, Chair of the IPCC, Abdalah Mokssit, Secretary thank MetadataWorks. Finally, we gratefully acknowledge funding
of the IPCC, and the staff of the IPCC Secretariat: Ermira Fida, Jonathan
x
Preface
support from the Governments of the United Kingdom and Germany, We thank our past WGI TSU team members: Wilfran Moufouma-Okia,
without which data archival at the DDC would not have been possible. Roz Pidcock, and Rodrigro Manzanas. We also thank the contributions
Preface
of Margot Eyraud, Evéa Piedagnel, Mathilde Mousson and Felix
A special thanks goes to the visual design team of the Summary Chavelli, who joined the TSU as interns.
for Policymakers: Tom Johansen and Angela Morelli of Information
Design Lab and Jordan Harold and Irene Lorenzoni of the Tyndall We wish to express our sincere recognition to all those who
Centre for Climate Change Research, as well as to Nigel Hawtin for contributed to the WGI assessment given the implications of
graphical design support of the Report. We would like to thank Alisa undertaking this during the COVID-19 pandemic, all of whom have
Singer for creating the “Changing” artwork inspired by one of the worked from home under such challenging conditions.
scientific figures for the front cover of the Report.
Finally, on behalf of all the participants to this unprecedented
Our particular appreciation goes to the WGI TSU, whose tireless experience, we would like to thank colleagues, friends, and
dedication, professionalism and enthusiasm underpinned the families who have also been part of this intense journey for their
production of this Report. The Report could not have been prepared understanding and support.
without the commitment of members of the TSU, all new to the IPCC,
who rose to the unprecedented Sixth Assessment Report challenge This report shows that how much climate change we experience in
and were pivotal in all aspects of the preparation of the Report: Anna the future depends on our decisions now, and what to prepare for.
Pirani, Clotilde Péan, Sarah Connors, Yang Chen, Robin Matthews, We wish that this report is widely used to provide evidence-based
Melissa Gomis, Sophie Berger, Leah Goldfarb, Rong Yu, Baiquan Zhou, knowledge to inform decision-making, for teaching and training, and
Ozge Yelekci, Nada Caud, Katherine Leitzell, Tom Maycock, Mengtian to enhance climate literacy worldwide.
Huang, Elisabeth Lonnoy, Tim Waterfield and Diego Cammarano.
xi
Dedication
Dedication
Dedication
Sir John Hougthon
(30 December 1931 – 15 April 2020)
The Working Group I Contribution to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (IPCC) Climate
Change 2021: The Physical Science Basis is dedicated to the memory of Sir John Hougthon, who was one of the key figures in the
creation of the IPCC in 1988, and served as Chair and Co-Chair of Working Group I for the IPCC’s first three assessment reports from
1988 to 2002.
Sir John’s work was a major factor in the award of the Nobel Peace Prize to the IPCC in 2007, shared with former U.S. Vice-President
Al Gore. He contributed to the development of climate science and building international cooperation based upon climate research.
Sir John played a key role in ensuring a robust science-policy interface, used in the IPCC process, but his role in international scientific
research extended beyond the IPCC, for instance in contributing to the establishment of the World Climate Research Programme,
which he chaired from 1982 to 1984.
Sir John was a brilliant communicator among scientific colleagues, policymakers and the public at large, explaining the fact and threat
of climate change with clarity and directness.
xiii
Contents
SPM
L. Connors (France/United Kingdom), Erika Coppola (Italy), Faye Abigail Cruz (Philippines), Aïda Diongue-
Niang (Senegal), Francisco J. Doblas-Reyes (Spain), Hervé Douville (France), Fatima Driouech (Morocco),
Tamsin L. Edwards (United Kingdom), François Engelbrecht (South Africa), Veronika Eyring (Germany),
Erich Fischer (Switzerland), Gregory M. Flato (Canada), Piers Forster (United Kingdom), Baylor Fox-
Kemper (United States of America), Jan S. Fuglestvedt (Norway), John C. Fyfe (Canada), Nathan P. Gillett
(Canada), Melissa I. Gomis (France/Switzerland), Sergey K. Gulev (Russian Federation), José Manuel
Gutiérrez (Spain), Rafiq Hamdi (Belgium), Jordan Harold (United Kingdom), Mathias Hauser (Switzerland),
Ed Hawkins (United Kingdom), Helene T. Hewitt (United Kingdom), Tom Gabriel Johansen (Norway),
Christopher Jones (United Kingdom), Richard G. Jones (United Kingdom), Darrell S. Kaufman (United
States of America), Zbigniew Klimont (Austria/Poland), Robert E. Kopp (United States of America), Charles
Koven (United States of America), Gerhard Krinner (France/Germany, France), June-Yi Lee (Republic of
Korea), Irene Lorenzoni (United Kingdom/Italy), Jochem Marotzke (Germany), Valérie Masson-Delmotte
(France), Thomas K. Maycock (United States of America), Malte Meinshausen (Australia/Germany), Pedro
M.S. Monteiro (South Africa), Angela Morelli (Norway/Italy), Vaishali Naik (United States of America),
Dirk Notz (Germany), Friederike Otto (United Kingdom/Germany), Matthew D. Palmer (United Kingdom),
Izidine Pinto (South Africa/Mozambique), Anna Pirani (Italy), Gian-Kasper Plattner (Switzerland),
Krishnan Raghavan (India), Roshanka Ranasinghe (The Netherlands/Sri Lanka, Australia), Joeri Rogelj
(United Kingdom/Belgium), Maisa Rojas (Chile), Alex C. Ruane (United States of America), Jean-Baptiste
Sallée (France), Bjørn H. Samset (Norway), Sonia I. Seneviratne (Switzerland), Jana Sillmann (Norway/
Germany), Anna A. Sörensson (Argentina), Tannecia S. Stephenson (Jamaica), Trude Storelvmo (Norway),
Sophie Szopa (France), Peter W. Thorne (Ireland/United Kingdom), Blair Trewin (Australia), Robert Vautard
(France), Carolina Vera (Argentina), Noureddine Yassaa (Algeria), Sönke Zaehle (Germany), Panmao Zhai
(China), Xuebin Zhang (Canada), Kirsten Zickfeld (Canada/Germany)
Contributing Authors:
Krishna M. AchutaRao (India), Bhupesh Adhikary (Nepal), Edvin Aldrian (Indonesia), Kyle Armour (United
States of America), Govindasamy Bala (India/United States of America), Rondrotiana Barimalala (South
Africa/Madagascar), Nicolas Bellouin (United Kingdom/France), William Collins (United Kingdom),
William D. Collins (United States of America), Susanna Corti (Italy), Peter M. Cox (United Kingdom),
Frank J. Dentener (EU/The Netherlands), Claudine Dereczynski (Brazil), Alejandro Di Luca (Australia,
Canada/Argentina), Alessandro Dosio (Italy), Leah Goldfarb (France/United States of America), Irina
V. Gorodetskaya (Portugal/Belgium, Russian Federation), Pandora Hope (Australia), Mark Howden
(Australia), A.K.M Saiful Islam (Bangladesh), Yu Kosaka (Japan), James Kossin (United States of America),
Svitlana Krakovska (Ukraine), Chao Li (China), Jian Li (China), Thorsten Mauritsen (Germany/Denmark),
Sebastian Milinski (Germany), Seung-Ki Min (Republic of Korea), Thanh Ngo Duc (Vietnam), Andy
Reisinger (New Zealand), Lucas Ruiz (Argentina), Shubha Sathyendranath (United Kingdom/Canada,
Overseas Citizen of India), Aimée B. A. Slangen (The Netherlands), Chris Smith (United Kingdom), Izuru
Takayabu (Japan), Muhammad Irfan Tariq (Pakistan), Anne-Marie Treguier (France), Bart van den Hurk
(The Netherlands), Karina von Schuckmann (France/Germany), Cunde Xiao (China)
3
Summary for Policymakers
Introduction
This Summary for Policymakers (SPM) presents key findings of the Working Group I (WGI) contribution to the Intergovernmental
Panel on Climate Change (IPCC) Sixth Assessment Report (AR6)1 on the physical science basis of climate change. The report builds
upon the 2013 Working Group I contribution to the IPCC’s Fifth Assessment Report (AR5) and the 2018–2019 IPCC Special Reports2
of the AR6 cycle and incorporates subsequent new evidence from climate science.3
This SPM provides a high-level summary of the understanding of the current state of the climate, including how it is changing and the
role of human influence, the state of knowledge about possible climate futures, climate information relevant to regions and sectors,
and limiting human-induced climate change.
Based on scientific understanding, key findings can be formulated as statements of fact or associated with an assessed level of
confidence indicated using the IPCC calibrated language.4
SPM The scientific basis for each key finding is found in chapter sections of the main Report and in the integrated synthesis presented
in the Technical Summary (hereafter TS), and is indicated in curly brackets. The AR6 WGI Interactive Atlas facilitates exploration of
these key synthesis findings, and supporting climate change information, across the WGI reference regions.5
A.1 It is unequivocal that human influence has warmed the atmosphere, ocean and land. Widespread and rapid
changes in the atmosphere, ocean, cryosphere and biosphere have occurred.
{2.2, 2.3, Cross-Chapter Box 2.3, 3.3, 3.4, 3.5, 3.6, 3.8, 5.2, 5.3, 6.4, 7.3, 8.3, 9.2, 9.3, 9.5, 9.6, Cross-Chapter
Box 9.1} (Figure SPM.1, Figure SPM.2)
A.1.1 Observed increases in well-mixed greenhouse gas (GHG) concentrations since around 1750 are unequivocally caused
by human activities. Since 2011 (measurements reported in AR5), concentrations have continued to increase in the
atmosphere, reaching annual averages of 410 parts per million (ppm) for carbon dioxide (CO2), 1866 parts per billion
(ppb) for methane (CH4), and 332 ppb for nitrous oxide (N2O) in 2019.6 Land and ocean have taken up a near-constant
proportion (globally about 56% per year) of CO2 emissions from human activities over the past six decades, with regional
differences (high confidence).7
{2.2, 5.2, 7.3, TS.2.2, Box TS.5}
1 Decision IPCC/XLVI-2.
2 The three Special Reports are: Global Warming of 1.5°C: An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse
gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty (SR1.5);
Climate Change and Land: An IPCC Special Report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in
terrestrial ecosystems (SRCCL); IPCC Special Report on the Ocean and Cryosphere in a Changing Climate (SROCC).
3 The assessment covers scientific literature accepted for publication by 31 January 2021.
4 Each finding is grounded in an evaluation of underlying evidence and agreement. A level of confidence is expressed using five qualifiers: very low, low, medium, high and very high,
and typeset in italics, for example, medium confidence. The following terms have been used to indicate the assessed likelihood of an outcome or result: virtually certain 99–100%
probability; very likely 90–100%; likely 66–100%; about as likely as not 33–66%; unlikely 0–33%; very unlikely 0–10%; and exceptionally unlikely 0–1%. Additional terms
(extremely likely 95–100%; more likely than not >50–100%; and extremely unlikely 0–5%) are also used when appropriate. Assessed likelihood is typeset in italics, for example,
very likely. This is consistent with AR5. In this Report, unless stated otherwise, square brackets [x to y] are used to provide the assessed very likely range, or 90% interval.
5 The Interactive Atlas is available at https://interactive-atlas.ipcc.ch
6 Other GHG concentrations in 2019 were: perfluorocarbons (PFCs) – 109 parts per trillion (ppt) CF4 equivalent; sulphur hexafluoride (SF6) – 10 ppt; nitrogen trifluoride (NF3) – 2 ppt;
hydrofluorocarbons (HFCs) – 237 ppt HFC-134a equivalent; other Montreal Protocol gases (mainly chlorofluorocarbons (CFCs) and hydrochlorofluorocarbons (HCFCs)) – 1032 ppt
CFC-12 equivalent). Increases from 2011 are 19 ppm for CO2, 63 ppb for CH4 and 8 ppb for N2O.
7 Land and ocean are not substantial sinks for other GHGs.
4
Summary for Policymakers
A.1.2 Each of the last four decades has been successively warmer than any decade that preceded it since 1850. Global
surface temperature8 in the first two decades of the 21st century (2001–2020) was 0.99 [0.84 to 1.10] °C higher than
1850–1900.9 Global surface temperature was 1.09 [0.95 to 1.20] °C higher in 2011–2020 than 1850–1900, with larger
increases over land (1.59 [1.34 to 1.83] °C) than over the ocean (0.88 [0.68 to 1.01] °C). The estimated increase in
global surface temperature since AR5 is principally due to further warming since 2003–2012 (+0.19 [0.16 to 0.22] °C).
Additionally, methodological advances and new datasets contributed approximately 0.1°C to the updated estimate of
warming in AR6.10
{2.3, Cross-Chapter Box 2.3} (Figure SPM.1)
A.1.3 The likely range of total human-caused global surface temperature increase from 1850–1900 to 2010–201911 is 0.8°C to
1.3°C, with a best estimate of 1.07°C. It is likely that well-mixed GHGs contributed a warming of 1.0°C to 2.0°C, other
human drivers (principally aerosols) contributed a cooling of 0.0°C to 0.8°C, natural drivers changed global surface
temperature by –0.1°C to +0.1°C, and internal variability changed it by –0.2°C to +0.2°C. It is very likely that well-mixed
GHGs were the main driver12 of tropospheric warming since 1979 and extremely likely that human-caused stratospheric
ozone depletion was the main driver of cooling of the lower stratosphere between 1979 and the mid-1990s. SPM
{3.3, 6.4, 7.3, TS.2.3, Cross-Section Box TS.1} (Figure SPM.2)
A.1.4 Globally averaged precipitation over land has likely increased since 1950, with a faster rate of increase since the 1980s
(medium confidence). It is likely that human influence contributed to the pattern of observed precipitation changes
since the mid-20th century and extremely likely that human influence contributed to the pattern of observed changes
in near-surface ocean salinity. Mid-latitude storm tracks have likely shifted poleward in both hemispheres since the
1980s, with marked seasonality in trends (medium confidence). For the Southern Hemisphere, human influence very likely
contributed to the poleward shift of the closely related extratropical jet in austral summer.
{2.3, 3.3, 8.3, 9.2, TS.2.3, TS.2.4, Box TS.6}
A.1.5 Human influence is very likely the main driver of the global retreat of glaciers since the 1990s and the decrease in Arctic
sea ice area between 1979–1988 and 2010–2019 (decreases of about 40% in September and about 10% in March). There
has been no significant trend in Antarctic sea ice area from 1979 to 2020 due to regionally opposing trends and large
internal variability. Human influence very likely contributed to the decrease in Northern Hemisphere spring snow cover
since 1950. It is very likely that human influence has contributed to the observed surface melting of the Greenland Ice
Sheet over the past two decades, but there is only limited evidence, with medium agreement, of human influence on the
Antarctic Ice Sheet mass loss.
{2.3, 3.4, 8.3, 9.3, 9.5, TS.2.5}
A.1.6 It is virtually certain that the global upper ocean (0–700 m) has warmed since the 1970s and extremely likely that human
influence is the main driver. It is virtually certain that human-caused CO2 emissions are the main driver of current global
acidification of the surface open ocean. There is high confidence that oxygen levels have dropped in many upper ocean
regions since the mid-20th century and medium confidence that human influence contributed to this drop.
{2.3, 3.5, 3.6, 5.3, 9.2, TS.2.4}
A.1.7 Global mean sea level increased by 0.20 [0.15 to 0.25] m between 1901 and 2018. The average rate of sea level rise was
1.3 [0.6 to 2.1] mm yr–1 between 1901 and 1971, increasing to 1.9 [0.8 to 2.9] mm yr–1 between 1971 and 2006, and
further increasing to 3.7 [3.2 to 4.2] mm yr–1 between 2006 and 2018 (high confidence). Human influence was very likely
the main driver of these increases since at least 1971.
{2.3, 3.5, 9.6, Cross-Chapter Box 9.1, Box TS.4}
8 The term ‘global surface temperature’ is used in reference to both global mean surface temperature and global surface air temperature throughout this SPM. Changes in these
quantities are assessed with high confidence to differ by at most 10% from one another, but conflicting lines of evidence lead to low confidence in the sign (direction) of any
difference in long-term trend. {Cross-Section Box TS.1}
9 The period 1850–1900 represents the earliest period of sufficiently globally complete observations to estimate global surface temperature and, consistent with AR5 and SR1.5, is
used as an approximation for pre-industrial conditions.
10 Since AR5, methodological advances and new datasets have provided a more complete spatial representation of changes in surface temperature, including in the Arctic. These
and other improvements have also increased the estimate of global surface temperature change by approximately 0.1°C, but this increase does not represent additional physical
warming since AR5.
11 The period distinction with A.1.2 arises because the attribution studies consider this slightly earlier period. The observed warming to 2010–2019 is 1.06 [0.88 to 1.21] °C.
12 Throughout this SPM, ‘main driver’ means responsible for more than 50% of the change.
5
Summary for Policymakers
A.1.8 Changes in the land biosphere since 1970 are consistent with global warming: climate zones have shifted poleward in
both hemispheres, and the growing season has on average lengthened by up to two days per decade since the 1950s
in the Northern Hemisphere extratropics (high confidence).
{2.3, TS.2.6}
0.5 0.5
0.2 simulated
natural only
0.0 0.0 (solar &
volcanic)
reconstructed
–0.5 –0.5
–1
1 500 1000 1500 1850 2020 1850 1900 1950 2000 2020
Figure SPM.1 | History of global temperature change and causes of recent warming
Panel (a) Changes in global surface temperature reconstructed from paleoclimate archives (solid grey line, years 1–2000) and from direct
observations (solid black line, 1850–2020), both relative to 1850–1900 and decadally averaged. The vertical bar on the left shows the estimated temperature
(very likely range) during the warmest multi-century period in at least the last 100,000 years, which occurred around 6500 years ago during the current interglacial
period (Holocene). The Last Interglacial, around 125,000 years ago, is the next most recent candidate for a period of higher temperature. These past warm periods
were caused by slow (multi-millennial) orbital variations. The grey shading with white diagonal lines shows the very likely ranges for the temperature reconstructions.
Panel (b) Changes in global surface temperature over the past 170 years (black line) relative to 1850–1900 and annually averaged, compared to
Coupled Model Intercomparison Project Phase 6 (CMIP6) climate model simulations (see Box SPM.1) of the temperature response to both human and natural
drivers (brown) and to only natural drivers (solar and volcanic activity, green). Solid coloured lines show the multi-model average, and coloured shades show the
very likely range of simulations. (See Figure SPM.2 for the assessed contributions to warming).
{2.3.1; Cross-Chapter Box 2.3; 3.3; TS.2.2; Cross-Section Box TS.1, Figure 1a}
6
Summary for Policymakers
Internal variability
Carbon dioxide
Methane
Nitrous oxide
Halogenated gases
Nitrogen oxides
Sulphur dioxide
Organic carbon
Ammonia
Black carbon
and irrigation
Land-use reflectance
Aviation contrails
Mainly contribute to Mainly contribute to
changes in changes in
non-CO₂ greenhouse gases anthropogenic aerosols
7
Summary for Policymakers
A.2 The scale of recent changes across the climate system as a whole – and the present state of many aspects of
the climate system – are unprecedented over many centuries to many thousands of years.
{2.2, 2.3, Cross-Chapter Box 2.1, 5.1} (Figure SPM.1)
A.2.1 In 2019, atmospheric CO2 concentrations were higher than at any time in at least 2 million years (high confidence), and
concentrations of CH4 and N2O were higher than at any time in at least 800,000 years (very high confidence). Since 1750,
increases in CO2 (47%) and CH4 (156%) concentrations far exceed – and increases in N2O (23%) are similar to – the natural
multi-millennial changes between glacial and interglacial periods over at least the past 800,000 years (very high confidence).
{2.2, 5.1, TS.2.2}
A.2.2 Global surface temperature has increased faster since 1970 than in any other 50-year period over at least the last 2000
years (high confidence). Temperatures during the most recent decade (2011–2020) exceed those of the most recent
multi-century warm period, around 6500 years ago13 [0.2°C to 1°C relative to 1850–1900] (medium confidence). Prior
to that, the next most recent warm period was about 125,000 years ago, when the multi-century temperature [0.5°C to
SPM 1.5°C relative to 1850–1900] overlaps the observations of the most recent decade (medium confidence).
{2.3, Cross-Chapter Box 2.1, Cross-Section Box TS.1} (Figure SPM.1)
A.2.3 In 2011–2020, annual average Arctic sea ice area reached its lowest level since at least 1850 (high confidence). Late
summer Arctic sea ice area was smaller than at any time in at least the past 1000 years (medium confidence). The global
nature of glacier retreat since the 1950s, with almost all of the world’s glaciers retreating synchronously, is unprecedented
in at least the last 2000 years (medium confidence).
{2.3, TS.2.5}
A.2.4 Global mean sea level has risen faster since 1900 than over any preceding century in at least the last 3000 years (high
confidence). The global ocean has warmed faster over the past century than since the end of the last deglacial transition
(around 11,000 years ago) (medium confidence). A long-term increase in surface open ocean pH occurred over the past
50 million years (high confidence). However, surface open ocean pH as low as recent decades is unusual in the last
2 million years (medium confidence).
{2.3, TS.2.4, Box TS.4}
A.3 Human-induced climate change is already affecting many weather and climate extremes in every region
across the globe. Evidence of observed changes in extremes such as heatwaves, heavy precipitation, droughts,
and tropical cyclones, and, in particular, their attribution to human influence, has strengthened since AR5.
{2.3, 3.3, 8.2, 8.3, 8.4, 8.5, 8.6, Box 8.1, Box 8.2, Box 9.2, 10.6, 11.2, 11.3, 11.4, 11.6, 11.7, 11.8, 11.9, 12.3}
(Figure SPM.3)
A.3.1 It is virtually certain that hot extremes (including heatwaves) have become more frequent and more intense across most
land regions since the 1950s, while cold extremes (including cold waves) have become less frequent and less severe, with
high confidence that human-induced climate change is the main driver14 of these changes. Some recent hot extremes
observed over the past decade would have been extremely unlikely to occur without human influence on the climate
system. Marine heatwaves have approximately doubled in frequency since the 1980s (high confidence), and human
influence has very likely contributed to most of them since at least 2006.
{Box 9.2, 11.2, 11.3, 11.9, TS.2.4, TS.2.6, Box TS.10} (Figure SPM.3)
A.3.2 The frequency and intensity of heavy precipitation events have increased since the 1950s over most land area for which
observational data are sufficient for trend analysis (high confidence), and human-induced climate change is likely the
main driver. Human-induced climate change has contributed to increases in agricultural and ecological droughts15 in some
regions due to increased land evapotranspiration16 (medium confidence).
{8.2, 8.3, 11.4, 11.6, 11.9, TS.2.6, Box TS.10} (Figure SPM.3)
13 As stated in section B.1, even under the very low emissions scenario SSP1-1.9, temperatures are assessed to remain elevated above those of the most recent decade until at least
2100 and therefore warmer than the century-scale period 6500 years ago.
14 As indicated in footnote 12, throughout this SPM, ‘main driver’ means responsible for more than 50% of the change.
15 Agricultural and ecological drought (depending on the affected biome): a period with abnormal soil moisture deficit, which results from combined shortage of precipitation
and excess evapotranspiration, and during the growing season impinges on crop production or ecosystem function in general (see Annex VII: Glossary). Observed changes in
meteorological droughts (precipitation deficits) and hydrological droughts (streamflow deficits) are distinct from those in agricultural and ecological droughts and are addressed in
the underlying AR6 material (Chapter 11).
16 The combined processes through which water is transferred to the atmosphere from open water and ice surfaces, bare soils and vegetation that make up the Earth’s surface (Glossary).
8
Summary for Policymakers
A.3.3 Decreases in global land monsoon precipitation17 from the 1950s to the 1980s are partly attributed to human-caused
Northern Hemisphere aerosol emissions, but increases since then have resulted from rising GHG concentrations and
decadal to multi-decadal internal variability (medium confidence). Over South Asia, East Asia and West Africa, increases
in monsoon precipitation due to warming from GHG emissions were counteracted by decreases in monsoon precipitation
due to cooling from human-caused aerosol emissions over the 20th century (high confidence). Increases in West African
monsoon precipitation since the 1980s are partly due to the growing influence of GHGs and reductions in the cooling
effect of human-caused aerosol emissions over Europe and North America (medium confidence).
{2.3, 3.3, 8.2, 8.3, 8.4, 8.5, 8.6, Box 8.1, Box 8.2, 10.6, Box TS.13}
A.3.4 It is likely that the global proportion of major (Category 3–5) tropical cyclone occurrence has increased over the last four
decades, and it is very likely that the latitude where tropical cyclones in the western North Pacific reach their peak intensity
has shifted northward; these changes cannot be explained by internal variability alone (medium confidence). There is low
confidence in long-term (multi-decadal to centennial) trends in the frequency of all-category tropical cyclones. Event
attribution studies and physical understanding indicate that human-induced climate change increases heavy precipitation
associated with tropical cyclones (high confidence), but data limitations inhibit clear detection of past trends on the SPM
global scale.
{8.2, 11.7, Box TS.10}
A.3.5 Human influence has likely increased the chance of compound extreme events18 since the 1950s. This includes increases in
the frequency of concurrent heatwaves and droughts on the global scale (high confidence), fire weather in some regions
of all inhabited continents (medium confidence), and compound flooding in some locations (medium confidence).
{11.6, 11.7, 11.8, 12.3, 12.4, TS.2.6, Table TS.5, Box TS.10}
17 The global monsoon is defined as the area in which the annual range (local summer minus local winter) of precipitation is greater than 2.5 mm day–1 (Glossary). Global land monsoon
precipitation refers to the mean precipitation over land areas within the global monsoon.
18 Compound extreme events are the combination of multiple drivers and/or hazards that contribute to societal or environmental risk (Glossary). Examples are concurrent heatwaves
and droughts, compound flooding (e.g., a storm surge in combination with extreme rainfall and/or river flow), compound fire weather conditions (i.e., a combination of hot, dry and
windy conditions), or concurrent extremes at different locations.
9
Summary for Policymakers
Climate change is already affecting every inhabited region across the globe,
with human influence contributing to many observed changes in weather
and climate extremes
(a) Synthesis of assessment of observed change in hot extremes and
confidence in human contribution to the observed changes in the world’s regions
Type of observed change
in hot extremes North
America GIC Europe
Increase (41) NWN NEN NEU RAR
Each hexagon corresponds IPCC AR6 WGI reference regions: North America: NWN (North-Western North America, NEN (North-Eastern North America), WNA
to one of the IPCC AR6 (Western North America), CNA (Central North America), ENA (Eastern North America), Central America: NCA (Northern Central America),
WGI reference regions SCA (Southern Central America), CAR (Caribbean), South America: NWS (North-Western South America), NSA (Northern South America), NES
(North-Eastern South America), SAM (South American Monsoon), SWS (South-Western South America), SES (South-Eastern South America),
North-Western SSA (Southern South America), Europe: GIC (Greenland/Iceland), NEU (Northern Europe), WCE (Western and Central Europe), EEU (Eastern
NWN Europe), MED (Mediterranean), Africa: MED (Mediterranean), SAH (Sahara), WAF (Western Africa), CAF (Central Africa), NEAF (North Eastern
North America
Africa), SEAF (South Eastern Africa), WSAF (West Southern Africa), ESAF (East Southern Africa), MDG (Madagascar), Asia: RAR (Russian
Arctic), WSB (West Siberia), ESB (East Siberia), RFE (Russian Far East), WCA (West Central Asia), ECA (East Central Asia), TIB (Tibetan Plateau),
EAS (East Asia), ARP (Arabian Peninsula), SAS (South Asia), SEA (South East Asia), Australasia: NAU (Northern Australia), CAU (Central
Australia), EAU (Eastern Australia), SAU (Southern Australia), NZ (New Zealand), Small Islands: CAR (Caribbean), PAC (Pacific Small Islands)
10
Summary for Policymakers
A.4 Improved knowledge of climate processes, paleoclimate evidence and the response of the climate system to
increasing radiative forcing gives a best estimate of equilibrium climate sensitivity of 3°C, with a narrower
range compared to AR5.
{2.2, 7.3, 7.4, 7.5, Box 7.2, 9.4, 9.5, 9.6, Cross-Chapter Box 9.1}
A.4.1 Human-caused radiative forcing of 2.72 [1.96 to 3.48] W m–2 in 2019 relative to 1750 has warmed the climate system. This
warming is mainly due to increased GHG concentrations, partly reduced by cooling due to increased aerosol concentrations.
The radiative forcing has increased by 0.43 W m–2 (19%) relative to AR5, of which 0.34 W m–2 is due to the increase in GHG
concentrations since 2011. The remainder is due to improved scientific understanding and changes in the assessment of
aerosol forcing, which include decreases in concentration and improvement in its calculation (high confidence).
{2.2, 7.3, TS.2.2, TS.3.1}
A.4.2 Human-caused net positive radiative forcing causes an accumulation of additional energy (heating) in the climate system,
partly reduced by increased energy loss to space in response to surface warming. The observed average rate of heating of
the climate system increased from 0.50 [0.32 to 0.69] W m–2 for the period 1971–200619 to 0.79 [0.52 to 1.06] W m–2 for
the period 2006–201820 (high confidence). Ocean warming accounted for 91% of the heating in the climate system, with
land warming, ice loss and atmospheric warming accounting for about 5%, 3% and 1%, respectively (high confidence).
{7.2, Box 7.2, TS.3.1}
A.4.3 Heating of the climate system has caused global mean sea level rise through ice loss on land and thermal expansion
from ocean warming. Thermal expansion explained 50% of sea level rise during 1971–2018, while ice loss from glaciers
contributed 22%, ice sheets 20% and changes in land-water storage 8%. The rate of ice-sheet loss increased by a factor
of four between 1992–1999 and 2010–2019. Together, ice-sheet and glacier mass loss were the dominant contributors to
global mean sea level rise during 2006–2018 (high confidence).
{9.4, 9.5, 9.6, Cross-Chapter Box 9.1}
A.4.4 The equilibrium climate sensitivity is an important quantity used to estimate how the climate responds to radiative
forcing. Based on multiple lines of evidence,21 the very likely range of equilibrium climate sensitivity is between 2°C (high
confidence) and 5°C (medium confidence). The AR6 assessed best estimate is 3°C with a likely range of 2.5°C to 4°C
(high confidence), compared to 1.5°C to 4.5°C in AR5, which did not provide a best estimate.
{7.4, 7.5, TS.3.2}
19 Cumulative energy increase of 282 [177 to 387] ZJ over 1971–2006 (1 ZJ = 1021 joules).
20 Cumulative energy increase of 152 [100 to 205] ZJ over 2006–2018.
21 Understanding of climate processes, the instrumental record, paleoclimates and model-based emergent constraints (Glossary).
11
Summary for Policymakers
Box SPM.1.1: This Report assesses the climate response to five illustrative scenarios that cover the range of possible future
development of anthropogenic drivers of climate change found in the literature. They start in 2015, and include scenarios22
with high and very high GHG emissions (SSP3-7.0 and SSP5-8.5) and CO2 emissions that roughly double from current
SPM levels by 2100 and 2050, respectively, scenarios with intermediate GHG emissions (SSP2-4.5) and CO2 emissions remaining
around current levels until the middle of the century, and scenarios with very low and low GHG emissions and CO2 emissions
declining to net zero around or after 2050, followed by varying levels of net negative CO2 emissions23 (SSP1-1.9 and
SSP1-2.6), as illustrated in Figure SPM.4. Emissions vary between scenarios depending on socio-economic assumptions,
levels of climate change mitigation and, for aerosols and non-methane ozone precursors, air pollution controls. Alternative
assumptions may result in similar emissions and climate responses, but the socio-economic assumptions and the feasibility
or likelihood of individual scenarios are not part of the assessment.
{1.6, Cross-Chapter Box 1.4, TS.1.3} (Figure SPM.4)
Box SPM.1.2: This Report assesses results from climate models participating in the Coupled Model Intercomparison Project
Phase 6 (CMIP6) of the World Climate Research Programme. These models include new and better representations of
physical, chemical and biological processes, as well as higher resolution, compared to climate models considered in previous
IPCC assessment reports. This has improved the simulation of the recent mean state of most large-scale indicators of climate
change and many other aspects across the climate system. Some differences from observations remain, for example in
regional precipitation patterns. The CMIP6 historical simulations assessed in this Report have an ensemble mean global
surface temperature change within 0.2°C of the observations over most of the historical period, and observed warming is
within the very likely range of the CMIP6 ensemble. However, some CMIP6 models simulate a warming that is either above
or below the assessed very likely range of observed warming.
{1.5, Cross-Chapter Box 2.2, 3.3, 3.8, TS.1.2, Cross-Section Box TS.1} (Figure SPM.1b, Figure SPM.2)
Box SPM.1.3: The CMIP6 models considered in this Report have a wider range of climate sensitivity than in CMIP5 models
and the AR6 assessed very likely range, which is based on multiple lines of evidence. These CMIP6 models also show
a higher average climate sensitivity than CMIP5 and the AR6 assessed best estimate. The higher CMIP6 climate sensitivity
values compared to CMIP5 can be traced to an amplifying cloud feedback that is larger in CMIP6 by about 20%.
{Box 7.1, 7.3, 7.4, 7.5, TS.3.2}
Box SPM.1.4: For the first time in an IPCC report, assessed future changes in global surface temperature, ocean warming
and sea level are constructed by combining multi-model projections with observational constraints based on past simulated
warming, as well as the AR6 assessment of climate sensitivity. For other quantities, such robust methods do not yet exist
to constrain the projections. Nevertheless, robust projected geographical patterns of many variables can be identified at
a given level of global warming, common to all scenarios considered and independent of timing when the global warming
level is reached.
{1.6, 4.3, 4.6, Box 4.1, 7.5, 9.2, 9.6, Cross-Chapter Box 11.1, Cross-Section Box TS.1}
22 Throughout this Report, the five illustrative scenarios are referred to as SSPx-y, where ‘SSPx’ refers to the Shared Socio-economic Pathway or ‘SSP’ describing the socio-economic
trends underlying the scenario, and ‘y’ refers to the approximate level of radiative forcing (in watts per square metre, or W m–2) resulting from the scenario in the year 2100.
A detailed comparison to scenarios used in earlier IPCC reports is provided in Section TS.1.3, and Sections 1.6 and 4.6. The SSPs that underlie the specific forcing scenarios used to
drive climate models are not assessed by WGI. Rather, the SSPx-y labelling ensures traceability to the underlying literature in which specific forcing pathways are used as input to the
climate models. IPCC is neutral with regard to the assumptions underlying the SSPs, which do not cover all possible scenarios. Alternative scenarios may be considered or developed.
23 Net negative CO2 emissions are reached when anthropogenic removals of CO2 exceed anthropogenic emissions (Glossary).
12
Summary for Policymakers
60 SSP5-8.5
10 SSP2-4.5
SSP1-2.6
SSP1-1.9
40 0
2015 2050 2100
(b) Contribution to global surface temperature increase from different emissions, with a dominant role of CO₂ emissions
Change in global surface temperature in 2081–2100 relative to 1850–1900 (ºC)
5 5 5 5 5
4 4 4 4 4
3 3 3 3 3
2 2 2 2 2
1 1 1 1 1
0 0 0 0 0
–1 –1 –1 –1 –1
Total CO₂ Non-CO₂ Aerosols Total CO₂ Non-CO₂ Aerosols Total CO₂ Non-CO₂ Aerosols Total CO₂ Non-CO₂ Aerosols Total CO₂ Non-CO₂ Aerosols
(observed) GHGs land use (observed) GHGs Land use (observed) GHGs Land use (observed) GHGs Land use (observed) GHGs Land use
Total warming (observed warming to date in darker shade), warming from CO₂, warming from non-CO₂ GHGs and cooling from changes in aerosols and land use
Figure SPM.4 | Future anthropogenic emissions of key drivers of climate change and warming contributions by groups of drivers for
the five illustrative scenarios used in this report
The five scenarios are SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0 and SSP5-8.5.
Panel (a) Annual anthropogenic (human-caused) emissions over the 2015–2100 period. Shown are emissions trajectories for carbon dioxide
(CO2) from all sectors (GtCO2/yr) (left graph) and for a subset of three key non-CO2 drivers considered in the scenarios: methane (CH4, MtCH4/yr, top-right
graph); nitrous oxide (N2O, MtN2O/yr, middle-right graph); and sulphur dioxide (SO2, MtSO2/yr, bottom-right graph, contributing to anthropogenic aerosols
in panel (b).
13
Summary for Policymakers
Panel (b) Warming contributions by groups of anthropogenic drivers and by scenario are shown as the change in global surface
temperature (°C) in 2081–2100 relative to 1850–1900, with indication of the observed warming to date. Bars and whiskers represent median values
and the very likely range, respectively. Within each scenario bar plot, the bars represent: total global warming (°C; ‘total’ bar) (see Table SPM.1); warming
contributions (°C) from changes in CO2 (‘CO2’ bar) and from non-CO2 greenhouse gases (GHGs; ‘non-CO2 GHGs’ bar: comprising well-mixed greenhouse
gases and ozone); and net cooling from other anthropogenic drivers (‘aerosols and land use’ bar: anthropogenic aerosols, changes in reflectance due to
land-use and irrigation changes, and contrails from aviation) (see Figure SPM.2, panel c, for the warming contributions to date for individual drivers). The
best estimate for observed warming in 2010–2019 relative to 1850–1900 (see Figure SPM.2, panel a) is indicated in the darker column in the ‘total’ bar.
Warming contributions in panel (b) are calculated as explained in Table SPM.1 for the total bar. For the other bars, the contribution by groups of drivers is
calculated with a physical climate emulator of global surface temperature that relies on climate sensitivity and radiative forcing assessments.
{Cross-Chapter Box 1.4; 4.6; Figure 4.35; 6.7; Figures 6.18, 6.22 and 6.24; 7.3; Cross-Chapter Box 7.1; Figure 7.7; Box TS.7; Figures TS.4 and TS.15}
B.1 Global surface temperature will continue to increase until at least mid-century under all emissions scenarios
considered. Global warming of 1.5°C and 2°C will be exceeded during the 21st century unless deep reductions
SPM in CO2 and other greenhouse gas emissions occur in the coming decades.
{2.3, Cross-Chapter Box 2.3, Cross-Chapter Box 2.4, 4.3, 4.4, 4.5} (Figure SPM.1, Figure SPM.4, Figure SPM.8,
Table SPM.1, Box SPM.1)
B.1.1 Compared to 1850–1900, global surface temperature averaged over 2081–2100 is very likely to be higher by 1.0°C to
1.8°C under the very low GHG emissions scenario considered (SSP1-1.9), by 2.1°C to 3.5°C in the intermediate GHG
emissions scenario (SSP2-4.5) and by 3.3°C to 5.7°C under the very high GHG emissions scenario (SSP5-8.5).24 The last
time global surface temperature was sustained at or above 2.5°C higher than 1850–1900 was over 3 million years ago
(medium confidence).
{2.3, Cross-Chapter Box 2.4, 4.3, 4.5, Box TS.2, Box TS.4, Cross-Section Box TS.1} (Table SPM.1)
Table SPM.1 | Changes in global surface temperature, which are assessed based on multiple lines of evidence, for selected 20-year time
periods and the five illustrative emissions scenarios considered. Temperature differences relative to the average global surface temperature of the
period 1850–1900 are reported in °C. This includes the revised assessment of observed historical warming for the AR5 reference period 1986–2005, which
in AR6 is higher by 0.08 [–0.01 to +0.12] °C than in AR5 (see footnote 10). Changes relative to the recent reference period 1995–2014 may be calculated
approximately by subtracting 0.85°C, the best estimate of the observed warming from 1850–1900 to 1995–2014.
{Cross-Chapter Box 2.3, 4.3, 4.4, Cross-Section Box TS.1}
SSP1-2.6 1.5 1.2 to 1.8 1.7 1.3 to 2.2 1.8 1.3 to 2.4
SSP2-4.5 1.5 1.2 to 1.8 2.0 1.6 to 2.5 2.7 2.1 to 3.5
SSP3-7.0 1.5 1.2 to 1.8 2.1 1.7 to 2.6 3.6 2.8 to 4.6
SSP5-8.5 1.6 1.3 to 1.9 2.4 1.9 to 3.0 4.4 3.3 to 5.7
B.1.2 Based on the assessment of multiple lines of evidence, global warming of 2°C, relative to 1850–1900, would be exceeded
during the 21st century under the high and very high GHG emissions scenarios considered in this report (SSP3-7.0 and
SSP5-8.5, respectively). Global warming of 2°C would extremely likely be exceeded in the intermediate GHG emissions
scenario (SSP2-4.5). Under the very low and low GHG emissions scenarios, global warming of 2°C is extremely unlikely
to be exceeded (SSP1-1.9) or unlikely to be exceeded (SSP1-2.6).25 Crossing the 2°C global warming level in the mid-
term period (2041–2060) is very likely to occur under the very high GHG emissions scenario (SSP5-8.5), likely to occur
under the high GHG emissions scenario (SSP3-7.0), and more likely than not to occur in the intermediate GHG emissions
scenario (SSP2-4.5).26
{4.3, Cross-Section Box TS.1} (Table SPM.1, Figure SPM.4, Box SPM.1)
24 Changes in global surface temperature are reported as running 20-year averages, unless stated otherwise.
25 SSP1-1.9 and SSP1-2.6 are scenarios that start in 2015 and have very low and low GHG emissions, respectively, and CO2 emissions declining to net zero around or after 2050,
followed by varying levels of net negative CO2 emissions.
26 Crossing is defined here as having the assessed global surface temperature change, averaged over a 20-year period, exceed a particular global warming level.
14
Summary for Policymakers
B.1.3 Global warming of 1.5°C relative to 1850–1900 would be exceeded during the 21st century under the intermediate, high
and very high GHG emissions scenarios considered in this report (SSP2-4.5, SSP3-7.0 and SSP5-8.5, respectively). Under
the five illustrative scenarios, in the near term (2021–2040), the 1.5°C global warming level is very likely to be exceeded
under the very high GHG emissions scenario (SSP5-8.5), likely to be exceeded under the intermediate and high GHG
emissions scenarios (SSP2-4.5 and SSP3-7.0), more likely than not to be exceeded under the low GHG emissions scenario
(SSP1-2.6) and more likely than not to be reached under the very low GHG emissions scenario (SSP1-1.9).27 Furthermore, for
the very low GHG emissions scenario (SSP1-1.9), it is more likely than not that global surface temperature would decline
back to below 1.5°C toward the end of the 21st century, with a temporary overshoot of no more than 0.1°C above 1.5°C
global warming.
{4.3, Cross-Section Box TS.1} (Table SPM.1, Figure SPM.4)
B.1.4 Global surface temperature in any single year can vary above or below the long-term human-induced trend, due to
substantial natural variability.28 The occurrence of individual years with global surface temperature change above a certain
level, for example 1.5°C or 2°C, relative to 1850–1900 does not imply that this global warming level has been reached.29
{Cross-Chapter Box 2.3, 4.3, 4.4, Box 4.1, Cross-Section Box TS.1} (Table SPM.1, Figure SPM.1, Figure SPM.8) SPM
B.2 Many changes in the climate system become larger in direct relation to increasing global warming. They
include increases in the frequency and intensity of hot extremes, marine heatwaves, heavy precipitation,
and, in some regions, agricultural and ecological droughts; an increase in the proportion of intense tropical
cyclones; and reductions in Arctic sea ice, snow cover and permafrost.
{4.3, 4.5, 4.6, 7.4, 8.2, 8.4, Box 8.2, 9.3, 9.5, Box 9.2, 11.1, 11.2, 11.3, 11.4, 11.6, 11.7, 11.9, Cross-Chapter Box
11.1, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.4, Atlas.5, Atlas.6, Atlas.7, Atlas.8, Atlas.9, Atlas.10, Atlas.11}
(Figure SPM.5, Figure SPM.6, Figure SPM.8)
B.2.1 It is virtually certain that the land surface will continue to warm more than the ocean surface (likely 1.4 to 1.7 times more).
It is virtually certain that the Arctic will continue to warm more than global surface temperature, with high confidence
above two times the rate of global warming.
{2.3, 4.3, 4.5, 4.6, 7.4, 11.1, 11.3, 11.9, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.4, Atlas.5, Atlas.6, Atlas.7, Atlas.8, Atlas.9,
Atlas.10, Atlas.11, Cross-Section Box TS.1, TS.2.6} (Figure SPM.5)
B.2.2 With every additional increment of global warming, changes in extremes continue to become larger. For example, every
additional 0.5°C of global warming causes clearly discernible increases in the intensity and frequency of hot extremes,
including heatwaves (very likely), and heavy precipitation (high confidence), as well as agricultural and ecological
droughts30 in some regions (high confidence). Discernible changes in intensity and frequency of meteorological droughts,
with more regions showing increases than decreases, are seen in some regions for every additional 0.5°C of global
warming (medium confidence). Increases in frequency and intensity of hydrological droughts become larger with
increasing global warming in some regions (medium confidence). There will be an increasing occurrence of some extreme
events unprecedented in the observational record with additional global warming, even at 1.5°C of global warming.
Projected percentage changes in frequency are larger for rarer events (high confidence).
{8.2, 11.2, 11.3, 11.4, 11.6, 11.9, Cross-Chapter Box 11.1, Cross-Chapter Box 12.1, TS.2.6} (Figure SPM.5, Figure SPM.6)
B.2.3 Some mid-latitude and semi-arid regions, and the South American Monsoon region, are projected to see the highest
increase in the temperature of the hottest days, at about 1.5 to 2 times the rate of global warming (high confidence). The
Arctic is projected to experience the highest increase in the temperature of the coldest days, at about three times the rate
of global warming (high confidence). With additional global warming, the frequency of marine heatwaves will continue
to increase (high confidence), particularly in the tropical ocean and the Arctic (medium confidence).
{Box 9.2, 11.1, 11.3, 11.9, Cross-Chapter Box 11.1, Cross-Chapter Box 12.1, 12.4, TS.2.4, TS.2.6} (Figure SPM.6)
27 The AR6 assessment of when a given global warming level is first exceeded benefits from the consideration of the illustrative scenarios, the multiple lines of evidence entering the
assessment of future global surface temperature response to radiative forcing, and the improved estimate of historical warming. The AR6 assessment is thus not directly comparable to
the SR1.5 SPM, which reported likely reaching 1.5°C global warming between 2030 and 2052, from a simple linear extrapolation of warming rates of the recent past. When considering
scenarios similar to SSP1-1.9 instead of linear extrapolation, the SR1.5 estimate of when 1.5°C global warming is first exceeded is close to the best estimate reported here.
28 Natural variability refers to climatic fluctuations that occur without any human influence, that is, internal variability combined with the response to external natural factors such as
volcanic eruptions, changes in solar activity and, on longer time scales, orbital effects and plate tectonics (Glossary).
29 The internal variability in any single year is estimated to be about ±0.25°C (5–95% range, high confidence).
30 Projected changes in agricultural and ecological droughts are primarily assessed based on total column soil moisture. See footnote 15 for definition and relation to precipitation
and evapotranspiration.
15
Summary for Policymakers
B.2.4 It is very likely that heavy precipitation events will intensify and become more frequent in most regions with additional global
warming. At the global scale, extreme daily precipitation events are projected to intensify by about 7% for each 1°C of global
warming (high confidence). The proportion of intense tropical cyclones (Category 4–5) and peak wind speeds of the most
intense tropical cyclones are projected to increase at the global scale with increasing global warming (high confidence).
{8.2, 11.4, 11.7, 11.9, Cross-Chapter Box 11.1, Box TS.6, TS.4.3.1} (Figure SPM.5, Figure SPM.6)
B.2.5 Additional warming is projected to further amplify permafrost thawing and loss of seasonal snow cover, of land ice and of
Arctic sea ice (high confidence). The Arctic is likely to be practically sea ice-free in September31 at least once before 2050
under the five illustrative scenarios considered in this report, with more frequent occurrences for higher warming levels.
There is low confidence in the projected decrease of Antarctic sea ice.
{4.3, 4.5, 7.4, 8.2, 8.4, Box 8.2, 9.3, 9.5, 12.4, Cross-Chapter Box 12.1, Atlas.5, Atlas.6, Atlas.8, Atlas.9, Atlas.11, TS.2.5}
(Figure SPM.8)
SPM
With every increment of global warming, changes get larger
in regional mean temperature, precipitation and soil moisture
(b) Annual mean temperature change (°C) Across warming levels, land areas warm more than ocean areas, and the
Arctic and Antarctica warm more than the tropics.
relative to 1850–1900
Simulated change at 1.5°C global warming Simulated change at 2°C global warming Simulated change at 4°C global warming
31 Monthly average sea ice area of less than 1 million km2, which is about 15% of the average September sea ice area observed in 1979–1988.
16
Summary for Policymakers
l a.
(c) Annual mean precipitation change (%) Precipitation is projected to increase over high latitudes, the equatorial
Pacific and parts of the monsoon regions, but decrease over parts of the
relative to 1850–1900 subtropics and in limited areas of the tropics.
Simulated change at 1.5°C global warming Simulated change at 2°C global warming Simulated change at 4°C global warming
(d) Annual mean total column soil Across warming levels, changes in soil moisture largely follow changes in
precipitation but also show some differences due to the influence of
moisture change (standard deviation) evapotranspiration.
Simulated change at 1.5°C global warming Simulated change at 2°C global warming Simulated change at 4°C global warming
Figure SPM.5 | Changes in annual mean surface temperature, precipitation, and soil moisture
Panel (a) Comparison of observed and simulated annual mean surface temperature change. The left map shows the observed changes in annual
mean surface temperature in the period 1850–2020 per °C of global warming (°C). The local (i.e., grid point) observed annual mean surface temperature changes
are linearly regressed against the global surface temperature in the period 1850–2020. Observed temperature data are from Berkeley Earth, the dataset with
the largest coverage and highest horizontal resolution. Linear regression is applied to all years for which data at the corresponding grid point is available. The
regression method was used to take into account the complete observational time series and thereby reduce the role of internal variability at the grid point level.
White indicates areas where time coverage was 100 years or less and thereby too short to calculate a reliable linear regression. The right map is based on model
simulations and shows change in annual multi-model mean simulated temperatures at a global warming level of 1°C (20-year mean global surface temperature
change relative to 1850–1900). The triangles at each end of the colour bar indicate out-of-bound values, that is, values above or below the given limits.
Panel (b) Simulated annual mean temperature change (°C), panel (c) precipitation change (%), and panel (d) total column soil moisture change
(standard deviation of interannual variability) at global warming levels of 1.5°C, 2°C and 4°C (20-year mean global surface temperature change relative
to 1850–1900). Simulated changes correspond to Coupled Model Intercomparison Project Phase 6 (CMIP6) multi-model mean change (median change for soil
moisture) at the corresponding global warming level, that is, the same method as for the right map in panel (a).
In panel (c), high positive percentage changes in dry regions may correspond to small absolute changes. In panel (d), the unit is the standard deviation
of interannual variability in soil moisture during 1850–1900. Standard deviation is a widely used metric in characterizing drought severity. A projected
reduction in mean soil moisture by one standard deviation corresponds to soil moisture conditions typical of droughts that occurred about once every six years
during 1850–1900. In panel (d), large changes in dry regions with little interannual variability in the baseline conditions can correspond to small absolute
change. The triangles at each end of the colour bars indicate out-of-bound values, that is, values above or below the given limits. Results from all models
reaching the corresponding warming level in any of the five illustrative scenarios (SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0 and SSP5-8.5) are averaged.
Maps of annual mean temperature and precipitation changes at a global warming level of 3°C are available in Figure 4.31 and Figure 4.32 in Section 4.6.
Corresponding maps of panels (b), (c) and (d), including hatching to indicate the level of model agreement at grid-cell level, are found in Figures 4.31, 4.32 and
11.19, respectively; as highlighted in Cross-Chapter Box Atlas.1, grid-cell level hatching is not informative for larger spatial scales (e.g., over AR6 reference regions)
where the aggregated signals are less affected by small-scale variability, leading to an increase in robustness.
{Figure 1.14, 4.6.1, Cross-Chapter Box 11.1, Cross-Chapter Box Atlas.1, TS.1.3.2, Figures TS.3 and TS.5}
17
Summary for Policymakers
1850–1900 Present 1°C 1.5°C 2°C 4°C 1850–1900 Present 1°C 1.5°C 2°C 4°C
SPM
Once now likely will likely will likely will likely Once now likely will likely will likely will likely
occurs occur occur occur occurs occur occur occur
2.8 times 4.1 times 5.6 times 9.4 times 4.8 times 8.6 times 13.9 times 39.2 times
(1.8–3.2) (2.8–4.7) (3.8–6.0) (8.3–9.6) (2.3–6.4) (4.3–10.7) (6.9–16.6) (27.0–41.4)
+6°C +6°C
INTENSITY increase
INTENSITY increase
+5°C +5°C
+4°C +4°C
+3°C +3°C
+2°C +2°C
+1°C +1°C
0°C 0°C
+1.2°C +1.9°C +2.6°C +5.1°C +1.2°C +2.0°C +2.7°C +5.3°C
hotter hotter hotter hotter hotter hotter hotter hotter
Heavy precipitation over land Agricultural & ecological droughts in drying regions
10-year event 10-year event
Frequency and increase in intensity of heavy 1-day Frequency and increase in intensity of an agricultural and ecological
precipitation event that occurred once in 10 years on drought event that occurred once in 10 years on average across
average in a climate without human influence drying regions in a climate without human influence
1850–1900 Present 1°C 1.5°C 2°C 4°C 1850–1900 Present 1°C 1.5°C 2°C 4°C
FREQUENCY per 10 years
FREQUENCY per 10 years
Once now likely will likely will likely will likely Once now likely will likely will likely will likely
occurs occur occur occur occurs occur occur occur
1.3 times 1.5 times 1.7 times 2.7 times 1.7 times 2.0 times 2.4 times 4.1 times
(1.2–1.4) (1.4–1.7) (1.6–2.0) (2.3–3.6) (0.7–4.1) (1.0–5.1) (1.3–5.8) (1.7–7.2)
+40% +2 sd
INTENSITY increase
INTENSITY increase
+30%
+20% +1 sd
+10%
0% 0 sd
+6.7% +10.5% +14.0% +30.2% +0.3 sd +0.5 sd +0.6 sd +1.0 sd
wetter wetter wetter wetter drier drier drier drier
Figure SPM.6 | Projected changes in the intensity and frequency of hot temperature extremes over land, extreme precipitation over land,
and agricultural and ecological droughts in drying regions
Projected changes are shown at global warming levels of 1°C, 1.5°C, 2°C, and 4°C and are relative to 1850–1900,9 representing a climate without human
influence. The figure depicts frequencies and increases in intensity of 10- or 50-year extreme events from the base period (1850–1900) under different global
warming levels.
Hot temperature extremes are defined as the daily maximum temperatures over land that were exceeded on average once in a decade (10-year event) or once
in 50 years (50-year event) during the 1850–1900 reference period. Extreme precipitation events are defined as the daily precipitation amount over land that
18
Summary for Policymakers
was exceeded on average once in a decade during the 1850–1900 reference period. Agricultural and ecological drought events are defined as the annual
average of total column soil moisture below the 10th percentile of the 1850–1900 base period. These extremes are defined on model grid box scale. For hot
temperature extremes and extreme precipitation, results are shown for the global land. For agricultural and ecological drought, results are shown for drying regions
only, which correspond to the AR6 regions in which there is at least medium confidence in a projected increase in agricultural and ecological droughts at the 2°C
warming level compared to the 1850–1900 base period in the Coupled Model Intercomparison Project Phase 6 (CMIP6). These regions include Western North
America, Central North America, Northern Central America, Southern Central America, Caribbean, Northern South America, North-Eastern South America, South
American Monsoon, South-Western South America, Southern South America, Western and Central Europe, Mediterranean, West Southern Africa, East Southern
Africa, Madagascar, Eastern Australia, and Southern Australia (Caribbean is not included in the calculation of the figure because of the too-small number of full land
grid cells). The non-drying regions do not show an overall increase or decrease in drought severity. Projections of changes in agricultural and ecological droughts
in the CMIP Phase 5 (CMIP5) multi-model ensemble differ from those in CMIP6 in some regions, including in parts of Africa and Asia. Assessments of projected
changes in meteorological and hydrological droughts are provided in Chapter 11.
In the ‘frequency’ section, each year is represented by a dot. The dark dots indicate years in which the extreme threshold is exceeded, while light dots are years
when the threshold is not exceeded. Values correspond to the medians (in bold) and their respective 5–95% range based on the multi-model ensemble from
simulations of CMIP6 under different Shared Socio-economic Pathway scenarios. For consistency, the number of dark dots is based on the rounded-up median.
In the ‘intensity’ section, medians and their 5–95% range, also based on the multi-model ensemble from simulations of CMIP6, are displayed as dark and
light bars, respectively. Changes in the intensity of hot temperature extremes and extreme precipitation are expressed as degree Celsius and percentage. As for
agricultural and ecological drought, intensity changes are expressed as fractions of standard deviation of annual soil moisture.
{11.1; 11.3; 11.4; 11.6; 11.9; Figures 11.12, 11.15, 11.6, 11.7, and 11.18} SPM
B.3 Continued global warming is projected to further intensify the global water cycle, including its variability,
global monsoon precipitation and the severity of wet and dry events.
{4.3, 4.4, 4.5, 4.6, 8.2, 8.3, 8.4, 8.5, Box 8.2, 11.4, 11.6, 11.9, 12.4, Atlas.3} (Figure SPM.5, Figure SPM.6)
B.3.1 There is strengthened evidence since AR5 that the global water cycle will continue to intensify as global temperatures
rise (high confidence), with precipitation and surface water flows projected to become more variable over most land
regions within seasons (high confidence) and from year to year (medium confidence). The average annual global land
precipitation is projected to increase by 0–5% under the very low GHG emissions scenario (SSP1-1.9), 1.5–8% for the
intermediate GHG emissions scenario (SSP2-4.5) and 1–13% under the very high GHG emissions scenario (SSP5-8.5) by
2081–2100 relative to 1995–2014 (likely ranges). Precipitation is projected to increase over high latitudes, the equatorial
Pacific and parts of the monsoon regions, but decrease over parts of the subtropics and limited areas in the tropics
in SSP2-4.5, SSP3-7.0 and SSP5-8.5 (very likely). The portion of the global land experiencing detectable increases or
decreases in seasonal mean precipitation is projected to increase (medium confidence). There is high confidence in an
earlier onset of spring snowmelt, with higher peak flows at the expense of summer flows in snow-dominated regions
globally.
{4.3, 4.5, 4.6, 8.2, 8.4, Atlas.3, TS.2.6, TS.4.3, Box TS.6} (Figure SPM.5)
B.3.2 A warmer climate will intensify very wet and very dry weather and climate events and seasons, with implications for
flooding or drought (high confidence), but the location and frequency of these events depend on projected changes in
regional atmospheric circulation, including monsoons and mid-latitude storm tracks. It is very likely that rainfall variability
related to the El Niño–Southern Oscillation is projected to be amplified by the second half of the 21st century in the
SSP2-4.5, SSP3-7.0 and SSP5-8.5 scenarios.
{4.3, 4.5, 4.6, 8.2, 8.4, 8.5, 11.4, 11.6, 11.9, 12.4, TS.2.6, TS.4.2, Box TS.6} (Figure SPM.5, Figure SPM.6)
B.3.3 Monsoon precipitation is projected to increase in the mid- to long term at the global scale, particularly over South and
South East Asia, East Asia and West Africa apart from the far west Sahel (high confidence). The monsoon season is
projected to have a delayed onset over North and South America and West Africa (high confidence) and a delayed retreat
over West Africa (medium confidence).
{4.4, 4.5, 8.2, 8.3, 8.4, Box 8.2, Box TS.13}
B.3.4 A projected southward shift and intensification of Southern Hemisphere summer mid-latitude storm tracks and associated
precipitation is likely in the long term under high GHG emissions scenarios (SSP3-7.0, SSP5-8.5), but in the near term
the effect of stratospheric ozone recovery counteracts these changes (high confidence). There is medium confidence in
a continued poleward shift of storms and their precipitation in the North Pacific, while there is low confidence in projected
changes in the North Atlantic storm tracks.
{4.4, 4.5, 8.4, TS.2.3, TS.4.2}
B.4 Under scenarios with increasing CO2 emissions, the ocean and land carbon sinks are projected to be less
effective at slowing the accumulation of CO2 in the atmosphere.
{4.3, 5.2, 5.4, 5.5, 5.6} (Figure SPM.7)
19
Summary for Policymakers
B.4.1 While natural land and ocean carbon sinks are projected to take up, in absolute terms, a progressively larger amount
of CO2 under higher compared to lower CO2 emissions scenarios, they become less effective, that is, the proportion of
emissions taken up by land and ocean decrease with increasing cumulative CO2 emissions. This is projected to result in
a higher proportion of emitted CO2 remaining in the atmosphere (high confidence).
{5.2, 5.4, Box TS.5} (Figure SPM.7)
B.4.2 Based on model projections, under the intermediate GHG emissions scenario that stabilizes atmospheric CO2 concentrations
this century (SSP2-4.5), the rates of CO2 taken up by the land and ocean are projected to decrease in the second half of
the 21st century (high confidence). Under the very low and low GHG emissions scenarios (SSP1-1.9, SSP1-2.6), where CO2
concentrations peak and decline during the 21st century, the land and ocean begin to take up less carbon in response
to declining atmospheric CO2 concentrations (high confidence) and turn into a weak net source by 2100 under SSP1-1.9
(medium confidence). It is very unlikely that the combined global land and ocean sink will turn into a source by 2100
under scenarios without net negative emissions (SSP2-4.5, SSP3-7.0, SSP5-8.5).32
{4.3, 5.4, 5.5, 5.6, Box TS.5, TS.3.3}
SPM
B.4.3 The magnitude of feedbacks between climate change and the carbon cycle becomes larger but also more uncertain
in high CO2 emissions scenarios (very high confidence). However, climate model projections show that the uncertainties in
atmospheric CO2 concentrations by 2100 are dominated by the differences between emissions scenarios (high confidence).
Additional ecosystem responses to warming not yet fully included in climate models, such as CO2 and CH4 fluxes from
wetlands, permafrost thaw and wildfires, would further increase concentrations of these gases in the atmosphere
(high confidence).
{5.4, Box TS.5, TS.3.2}
The proportion of CO₂ emissions taken up by land and ocean carbon sinks
is smaller in scenarios with higher cumulative CO₂ emissions
Total cumulative CO₂ emissions taken up by land and ocean (colours) and remaining in the atmosphere (grey)
under the five illustrative scenarios from 1850 to 2100
GtCO₂
12,000
For scenarios with
higher cumulative
10,000 CO₂ emissions…
8000
ATMOSPHERE
6000
ATMOSPHERE …the amount of CO₂ emissions
taken up by land and ocean
4000 ATMOSPHERE carbon sinks is larger,
OCEAN
but more of the emitted
ATMOSPHERE
ATMOSPHERE
OCEAN CO₂ remains in the
2000 OCEAN
OCEAN atmosphere…
OCEAN
ND ND ND ND ND
L
E E E E E
A
OC OC OC OC OC
Figure SPM.7 | Cumulative anthropogenic CO2 emissions taken up by land and ocean sinks by 2100 under the five illustrative scenarios
The cumulative anthropogenic (human-caused) carbon dioxide (CO2) emissions taken up by the land and ocean sinks under the five illustrative scenarios (SSP1-1.9,
SSP1-2.6, SSP2-4.5, SSP3-7.0 and SSP5-8.5) are simulated from 1850 to 2100 by Coupled Model Intercomparison Project Phase 6 (CMIP6) climate models in the
concentration-driven simulations. Land and ocean carbon sinks respond to past, current and future emissions; therefore, cumulative sinks from 1850 to 2100 are
presented here. During the historical period (1850–2019) the observed land and ocean sink took up 1430 GtCO2 (59% of the emissions).
32 These projected adjustments of carbon sinks to stabilization or decline of atmospheric CO2 are accounted for in calculations of remaining carbon budgets.
20
Summary for Policymakers
The bar chart illustrates the projected amount of cumulative anthropogenic CO2 emissions (GtCO2) between 1850 and 2100 remaining in the atmosphere (grey
part) and taken up by the land and ocean (coloured part) in the year 2100. The doughnut chart illustrates the proportion of the cumulative anthropogenic
CO2 emissions taken up by the land and ocean sinks and remaining in the atmosphere in the year 2100. Values in % indicate the proportion of the cumulative
anthropogenic CO2 emissions taken up by the combined land and ocean sinks in the year 2100. The overall anthropogenic carbon emissions are calculated by
adding the net global land-use emissions from the CMIP6 scenario database to the other sectoral emissions calculated from climate model runs with prescribed CO2
concentrations.33 Land and ocean CO2 uptake since 1850 is calculated from the net biome productivity on land, corrected for CO2 losses due to land-use change by
adding the land-use change emissions, and net ocean CO2 flux.
{5.2.1; Table 5.1; 5.4.5; Figure 5.25; Box TS.5; Box TS.5, Figure 1}
B.5 Many changes due to past and future greenhouse gas emissions are irreversible for centuries to millennia,
especially changes in the ocean, ice sheets and global sea level.
{2.3, Cross-Chapter Box 2.4, 4.3, 4.5, 4.7, 5.3, 9.2, 9.4, 9.5, 9.6, Box 9.4} (Figure SPM.8)
B.5.1 Past GHG emissions since 1750 have committed the global ocean to future warming (high confidence). Over the rest of
the 21st century, likely ocean warming ranges from 2–4 (SSP1-2.6) to 4–8 times (SSP5-8.5) the 1971–2018 change. Based
on multiple lines of evidence, upper ocean stratification (virtually certain), ocean acidification (virtually certain) and ocean SPM
deoxygenation (high confidence) will continue to increase in the 21st century, at rates dependent on future emissions.
Changes are irreversible on centennial to millennial time scales in global ocean temperature (very high confidence),
deep-ocean acidification (very high confidence) and deoxygenation (medium confidence).
{4.3, 4.5, 4.7, 5.3, 9.2, TS.2.4} (Figure SPM.8)
B.5.2 Mountain and polar glaciers are committed to continue melting for decades or centuries (very high confidence). Loss of
permafrost carbon following permafrost thaw is irreversible at centennial time scales (high confidence). Continued ice
loss over the 21st century is virtually certain for the Greenland Ice Sheet and likely for the Antarctic Ice Sheet. There is
high confidence that total ice loss from the Greenland Ice Sheet will increase with cumulative emissions. There is limited
evidence for low-likelihood, high-impact outcomes (resulting from ice-sheet instability processes characterized by deep
uncertainty and in some cases involving tipping points) that would strongly increase ice loss from the Antarctic Ice Sheet
for centuries under high GHG emissions scenarios.34
{4.3, 4.7, 5.4, 9.4, 9.5, Box 9.4, Box TS.1, TS.2.5}
B.5.3 It is virtually certain that global mean sea level will continue to rise over the 21st century. Relative to 1995–2014, the likely
global mean sea level rise by 2100 is 0.28–0.55 m under the very low GHG emissions scenario (SSP1-1.9); 0.32–0.62 m
under the low GHG emissions scenario (SSP1-2.6); 0.44–0.76 m under the intermediate GHG emissions scenario (SSP2-4.5);
and 0.63–1.01 m under the very high GHG emissions scenario (SSP5-8.5); and by 2150 is 0.37–0.86 m under the very
low scenario (SSP1-1.9); 0.46–0.99 m under the low scenario (SSP1-2.6); 0.66–1.33 m under the intermediate scenario
(SSP2-4.5); and 0.98–1.88 m under the very high scenario (SSP5-8.5) (medium confidence).35 Global mean sea level rise
above the likely range – approaching 2 m by 2100 and 5 m by 2150 under a very high GHG emissions scenario (SSP5-8.5)
(low confidence) – cannot be ruled out due to deep uncertainty in ice-sheet processes.
{4.3, 9.6, Box 9.4, Box TS.4} (Figure SPM.8)
B.5.4 In the longer term, sea level is committed to rise for centuries to millennia due to continuing deep-ocean warming and
ice-sheet melt and will remain elevated for thousands of years (high confidence). Over the next 2000 years, global mean
sea level will rise by about 2 to 3 m if warming is limited to 1.5°C, 2 to 6 m if limited to 2°C and 19 to 22 m with 5°C of
warming, and it will continue to rise over subsequent millennia (low confidence). Projections of multi-millennial global
mean sea level rise are consistent with reconstructed levels during past warm climate periods: likely 5–10 m higher than
today around 125,000 years ago, when global temperatures were very likely 0.5°C–1.5°C higher than 1850–1900; and very
likely 5–25 m higher roughly 3 million years ago, when global temperatures were 2.5°C–4°C higher (medium confidence).
{2.3, Cross-Chapter Box 2.4, 9.6, Box TS.2, Box TS.4, Box TS.9}
33 The other sectoral emissions are calculated as the residual of the net land and ocean CO2 uptake and the prescribed atmospheric CO2 concentration changes in the CMIP6
simulations. These calculated emissions are net emissions and do not separate gross anthropogenic emissions from removals, which are included implicitly.
34 Low-likelihood, high-impact outcomes are those whose probability of occurrence is low or not well known (as in the context of deep uncertainty) but whose potential impacts on
society and ecosystems could be high. A tipping point is a critical threshold beyond which a system reorganizes, often abruptly and/or irreversibly. (Glossary) {1.4, Cross-Chapter Box
1.3, 4.7}
35 To compare to the 1986–2005 baseline period used in AR5 and SROCC, add 0.03 m to the global mean sea level rise estimates. To compare to the 1900 baseline period used in
Figure SPM.8, add 0.16 m.
21
Summary for Policymakers
Human activities affect all the major climate system components, with
some responding over decades and others over centuries
(a) Global surface temperature change relative to 1850–1900 (e) Global mean sea
°C
level change in 2300
5
relative to 1900
SSP5-8.5
4 SSP3-7.0 Sea level rise greater than
15 m cannot be ruled out
3 SSP2-4.5
with high emissions
2 SSP1-2.6
SSP1-1.9
1
0 9m
–1
SPM 1950 2000 2015 2050 2100
7m
6
SSP1-1.9
2 SSP1-2.6
6m
Practically ice-free SSP2-4.5
SSP3-7.0
0 SSP5-8.5
1950 2000 2015 2050 2100
8.2
SSP5-8.5
8.1
SSP1-1.9
8.0 SSP1-2.6
7.9 SSP2-4.5
4m
ocean
7.8 acidification
SSP3-7.0
7.7
SSP5-8.5
7.6
1950 2000 2015 2050 2100 3m
2 2m
1.5
Low-likelihood, high-impact storyline,
including ice-sheet instability
processes, under SSP5-8.5
1 SSP5-8.5 1m
SSP3-7.0
SSP2-4.5
SSP1-2.6
0.5 SSP1-1.9
0 0m
1950 2000 2020 2050 2100 2300
Figure SPM.8 | Selected indicators of global climate change under the five illustrative scenarios used in this Report
The projections for each of the five scenarios are shown in colour. Shades represent uncertainty ranges – more detail is provided for each panel below. The black
curves represent the historical simulations (panels a, b, c) or the observations (panel d). Historical values are included in all graphs to provide context for the
projected future changes.
22
Summary for Policymakers
Panel (a) Global surface temperature changes in °C relative to 1850–1900. These changes were obtained by combining Coupled Model Intercomparison
Project Phase 6 (CMIP6) model simulations with observational constraints based on past simulated warming, as well as an updated assessment of equilibrium climate
sensitivity (see Box SPM.1). Changes relative to 1850–1900 based on 20-year averaging periods are calculated by adding 0.85°C (the observed global surface
temperature increase from 1850–1900 to 1995–2014) to simulated changes relative to 1995–2014. Very likely ranges are shown for SSP1-2.6 and SSP3-7.0.
Panel (b) September Arctic sea ice area in 106 km2 based on CMIP6 model simulations. Very likely ranges are shown for SSP1-2.6 and SSP3-7.0. The Arctic is
projected to be practically ice-free near mid-century under intermediate and high GHG emissions scenarios.
Panel (c) Global ocean surface pH (a measure of acidity) based on CMIP6 model simulations. Very likely ranges are shown for SSP1-2.6 and SSP3-7.0.
Panel (d) Global mean sea level change in metres, relative to 1900. The historical changes are observed (from tide gauges before 1992 and altimeters
afterwards), and the future changes are assessed consistently with observational constraints based on emulation of CMIP, ice-sheet, and glacier models. Likely
ranges are shown for SSP1-2.6 and SSP3-7.0. Only likely ranges are assessed for sea level changes due to difficulties in estimating the distribution of deeply
uncertain processes. The dashed curve indicates the potential impact of these deeply uncertain processes. It shows the 83rd percentile of SSP5-8.5 projections that
include low-likelihood, high-impact ice-sheet processes that cannot be ruled out; because of low confidence in projections of these processes, this curve does not
constitute part of a likely range. Changes relative to 1900 are calculated by adding 0.158 m (observed global mean sea level rise from 1900 to 1995–2014) to
simulated and observed changes relative to 1995–2014.
Panel (e) Global mean sea level change at 2300 in metres relative to 1900. Only SSP1-2.6 and SSP5-8.5 are projected at 2300, as simulations that extend
beyond 2100 for the other scenarios are too few for robust results. The 17th–83rd percentile ranges are shaded. The dashed arrow illustrates the 83rd percentile
of SSP5-8.5 projections that include low-likelihood, high-impact ice-sheet processes that cannot be ruled out.
SPM
Panels (b) and (c) are based on single simulations from each model, and so include a component of internal variability. Panels (a), (d) and (e) are based on long-term
averages, and hence the contributions from internal variability are small.
{4.3; Figures 4.2, 4.8, and 4.11; 9.6; Figure 9.27; Figures TS.8 and TS.11; Box TS.4, Figure 1}
C.1 Natural drivers and internal variability will modulate human-caused changes, especially at regional scales and
in the near term, with little effect on centennial global warming. These modulations are important to consider
in planning for the full range of possible changes.
{1.4, 2.2, 3.3, Cross-Chapter Box 3.1, 4.4, 4.6, Cross-Chapter Box 4.1, Box 7.2, 8.3, 8.5, 9.2, 10.3, 10.4, 10.6,
11.3, 12.5, Atlas.4, Atlas.5, Atlas.8, Atlas.9, Atlas.10, Atlas.11, Cross-Chapter Box Atlas.2}
C.1.1 The historical global surface temperature record highlights that decadal variability has both enhanced and masked
underlying human-caused long-term changes, and this variability will continue into the future (very high confidence). For
example, internal decadal variability and variations in solar and volcanic drivers partially masked human-caused surface
global warming during 1998–2012, with pronounced regional and seasonal signatures (high confidence). Nonetheless,
the heating of the climate system continued during this period, as reflected in both the continued warming of the global
ocean (very high confidence) and in the continued rise of hot extremes over land (medium confidence).
{1.4, 3.3, Cross-Chapter Box 3.1, 4.4, Box 7.2, 9.2, 11.3, Cross-Section Box TS.1} (Figure SPM.1)
C.1.2 Projected human-caused changes in mean climate and climatic impact-drivers (CIDs),36 including extremes, will be either
amplified or attenuated by internal variability (high confidence).37 Near-term cooling at any particular location with
respect to present climate could occur and would be consistent with the global surface temperature increase due to
human influence (high confidence).
{1.4, 4.4, 4.6, 10.4, 11.3, 12.5, Atlas.5, Atlas.10, Atlas.11, TS.4.2}
36 Climatic impact-drivers (CIDs) are physical climate system conditions (e.g., means, events, extremes) that affect an element of society or ecosystems. Depending on system tolerance,
CIDs and their changes can be detrimental, beneficial, neutral, or a mixture of each across interacting system elements and regions (Glossary). CID types include heat and cold, wet
and dry, wind, snow and ice, coastal and open ocean.
37 The main internal variability phenomena include El Niño–Southern Oscillation, Pacific Decadal Variability and Atlantic Multi-decadal Variability through their regional influence.
23
Summary for Policymakers
C.1.3 Internal variability has largely been responsible for the amplification and attenuation of the observed human-caused
decadal-to-multi-decadal mean precipitation changes in many land regions (high confidence). At global and regional
scales, near-term changes in monsoons will be dominated by the effects of internal variability (medium confidence).
In addition to the influence of internal variability, near-term projected changes in precipitation at global and regional
scales are uncertain because of model uncertainty and uncertainty in forcings from natural and anthropogenic aerosols
(medium confidence).
{1.4, 4.4, 8.3, 8.5, 10.3, 10.4, 10.5, 10.6, Atlas.4, Atlas.8, Atlas.9, Atlas.10, Atlas.11, Cross-Chapter Box Atlas.2, TS.4.2,
Box TS.6, Box TS.13}
C.1.4 Based on paleoclimate and historical evidence, it is likely that at least one large explosive volcanic eruption would occur
during the 21st century.38 Such an eruption would reduce global surface temperature and precipitation, especially over land,
for one to three years, alter the global monsoon circulation, modify extreme precipitation and change many CIDs (medium
confidence). If such an eruption occurs, this would therefore temporarily and partially mask human-caused climate change.
{2.2, 4.4, Cross-Chapter Box 4.1, 8.5, TS.2.1}
SPM
C.2 With further global warming, every region is projected to increasingly experience concurrent and multiple
changes in climatic impact-drivers. Changes in several climatic impact-drivers would be more widespread
at 2°C compared to 1.5°C global warming and even more widespread and/or pronounced for higher
warming levels.
{8.2, 9.3, 9.5, 9.6, Box 10.3, 11.3, 11.4, 11.5, 11.6, 11.7, 11.9, Box 11.3, Box 11.4, Cross-Chapter Box 11.1, 12.2,
12.3, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.4, Atlas.5, Atlas.6, Atlas.7, Atlas.8, Atlas.9, Atlas.10, Atlas.11}
(Table SPM.1, Figure SPM.9)
C.2.1 All regions39 are projected to experience further increases in hot climatic impact-drivers (CIDs) and decreases in cold
CIDs (high confidence). Further decreases are projected in permafrost; snow, glaciers and ice sheets; and lake and Arctic
sea ice (medium to high confidence).40 These changes would be larger at 2°C global warming or above than at 1.5°C
(high confidence). For example, extreme heat thresholds relevant to agriculture and health are projected to be exceeded
more frequently at higher global warming levels (high confidence).
{9.3, 9.5, 11.3, 11.9, Cross-Chapter Box 11.1, 12.3, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.4, Atlas.5, Atlas.6, Atlas.7,
Atlas.8, Atlas.9, Atlas.10, Atlas.11, TS.4.3} (Table SPM.1, Figure SPM.9)
C.2.2 At 1.5°C global warming, heavy precipitation and associated flooding are projected to intensify and be more frequent
in most regions in Africa and Asia (high confidence), North America (medium to high confidence)40 and Europe (medium
confidence). Also, more frequent and/or severe agricultural and ecological droughts are projected in a few regions in all
inhabited continents except Asia compared to 1850–1900 (medium confidence); increases in meteorological droughts are
also projected in a few regions (medium confidence). A small number of regions are projected to experience increases or
decreases in mean precipitation (medium confidence).
{11.4, 11.5, 11.6, 11.9, Atlas.4, Atlas.5, Atlas.7, Atlas.8, Atlas.9, Atlas.10, Atlas.11, TS.4.3} (Table SPM.1)
C.2.3 At 2°C global warming and above, the level of confidence in and the magnitude of the change in droughts and heavy
and mean precipitation increase compared to those at 1.5°C. Heavy precipitation and associated flooding events
are projected to become more intense and frequent in the Pacific Islands and across many regions of North America
and Europe (medium to high confidence).40 These changes are also seen in some regions in Australasia and Central and
South America (medium confidence). Several regions in Africa, South America and Europe are projected to experience an
increase in frequency and/or severity of agricultural and ecological droughts with medium to high confidence;40 increases
are also projected in Australasia, Central and North America, and the Caribbean with medium confidence. A small number
of regions in Africa, Australasia, Europe and North America are also projected to be affected by increases in hydrological
droughts, and several regions are projected to be affected by increases or decreases in meteorological droughts, with
more regions displaying an increase (medium confidence). Mean precipitation is projected to increase in all polar, northern
European and northern North American regions, most Asian regions and two regions of South America (high confidence).
{11.4, 11.6, 11.9, Cross-Chapter Box 11.1, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.5, Atlas.7, Atlas.8, Atlas.9, Atlas.11,
TS.4.3} (Table SPM.1, Figure SPM.5, Figure SPM.6, Figure SPM.9)
38 Based on 2500 year reconstructions, eruptions more negative than –1 W m–2 occur on average twice per century.
39 Regions here refer to the AR6 WGI reference regions used in this Report to summarize information in sub-continental and oceanic regions. Changes are compared to averages over
the last 20–40 years unless otherwise specified. {1.4, 12.4, Atlas.1}.
40 The specific level of confidence or likelihood depends on the region considered. Details can be found in the Technical Summary and the underlying Report.
24
Summary for Policymakers
C.2.4 More CIDs across more regions are projected to change at 2°C and above compared to 1.5°C global warming
(high confidence). Region-specific changes include intensification of tropical cyclones and/or extratropical storms
(medium confidence), increases in river floods (medium to high confidence),40 reductions in mean precipitation and
increases in aridity (medium to high confidence),40 and increases in fire weather (medium to high confidence).40 There
is low confidence in most regions in potential future changes in other CIDs, such as hail, ice storms, severe storms, dust
storms, heavy snowfall and landslides.
{11.7, 11.9, Cross-Chapter Box 11.1, 12.4, 12.5, Cross-Chapter Box 12.1, Atlas.4, Atlas.6, Atlas.7, Atlas.8, Atlas.10, TS.4.3.1,
TS.4.3.2, TS.5} (Table SPM.1, Figure SPM.9)
C.2.5 It is very likely to virtually certain40 that regional mean relative sea level rise will continue throughout the 21st century,
except in a few regions with substantial geologic land uplift rates. Approximately two-thirds of the global coastline has
a projected regional relative sea level rise within ±20% of the global mean increase (medium confidence). Due to relative
sea level rise, extreme sea level events that occurred once per century in the recent past are projected to occur at least
annually at more than half of all tide gauge locations by 2100 (high confidence). Relative sea level rise contributes to
increases in the frequency and severity of coastal flooding in low-lying areas and to coastal erosion along most sandy SPM
coasts (high confidence).
{9.6, 12.4, 12.5, Cross-Chapter Box 12.1, Box TS.4, TS.4.3} (Figure SPM.9)
C.2.6 Cities intensify human-induced warming locally, and further urbanization together with more frequent hot extremes will
increase the severity of heatwaves (very high confidence). Urbanization also increases mean and heavy precipitation
over and/or downwind of cities (medium confidence) and resulting runoff intensity (high confidence). In coastal cities,
the combination of more frequent extreme sea level events (due to sea level rise and storm surge) and extreme rainfall/
riverflow events will make flooding more probable (high confidence).
{8.2, Box 10.3, 11.3, 12.4, Box TS.14}
C.2.7 Many regions are projected to experience an increase in the probability of compound events with higher global warming
(high confidence). In particular, concurrent heatwaves and droughts are likely to become more frequent. Concurrent
extremes at multiple locations, including in crop-producing areas, become more frequent at 2°C and above compared to
1.5°C global warming (high confidence).
{11.8, Box 11.3, Box 11.4, 12.3, 12.4, Cross-Chapter Box 12.1, TS.4.3} (Table SPM.1)
25
Summary for Policymakers
Number of land & coastal regions (a) and open-ocean regions (b) where each climatic impact-driver (CID) is projected
to increase or decrease with high confidence (dark shade) or medium confidence (light shade)
(a) (b)
Heat and Cold Wet and Dry Wind Snow and Ice Other Coastal Open Ocean
SPM
Heavy precipitation and pluvial flood
Radiation at surface
Mean precipitation
Dissolved oxygen
Marine heatwave
Marine heatwave
Relative sea level
Snow avalanche
Tropical cyclone
Coastal erosion
Ocean salinity
Ocean acidity
Ocean acidity
NUMBER OF LAND
Extreme heat
Coastal flood
Fire weather
Permafrost
River flood
Cold spell
Landslide
Aridity
Frost
Hail
55
45
35
25
15 15
5 5
5 5
15 15
25
35
45
55
Regions with high confidence increase The height of the lighter shaded ‘envelope’ behind each bar Changes refer to a 20–30
Regions with medium confidence increase represents the maximum number of regions for which each year period centred around
CID is relevant. The envelope is symmetrical about the x-axis 2050 and/or consistent
Regions with high confidence decrease showing the maximum possible number of relevant regions with 2°C global warming
Regions with medium confidence decrease for CID increase (upper part) or decrease (lower part). compared to a similar
period within 1960–2014
or 1850–1900.
Figure SPM.9 | Synthesis of the number of AR6 WGI reference regions where climatic impact-drivers are projected to change
A total of 35 climatic impact-drivers (CIDs) grouped into seven types are shown: heat and cold; wet and dry; wind; snow and ice; coastal; open ocean; and other.
For each CID, the bar in the graph below displays the number of AR6 WGI reference regions where it is projected to change. The colours represent the direction
of change and the level of confidence in the change: purple indicates an increase while brown indicates a decrease; darker and lighter shades refer to high and
medium confidence, respectively. Lighter background colours represent the maximum number of regions for which each CID is broadly relevant.
Panel (a) shows the 30 CIDs relevant to the land and coastal regions, while panel (b) shows the five CIDs relevant to the open-ocean regions. Marine heatwaves
and ocean acidity are assessed for coastal ocean regions in panel (a) and for open-ocean regions in panel (b). Changes refer to a 20–30-year period centred around 2050
and/or consistent with 2°C global warming compared to a similar period within 1960–2014, except for hydrological drought and agricultural and ecological drought, which
is compared to 1850–1900. Definitions of the regions are provided in Sections 12.4 and Atlas.1 and the Interactive Atlas (see https://interactive-atlas.ipcc.ch/).
{11.9, 12.2, 12.4, Atlas.1, Table TS.5, Figures TS.22 and TS.25} (Table SPM.1)
26
Summary for Policymakers
C.3 Low-likelihood outcomes, such as ice-sheet collapse, abrupt ocean circulation changes, some compound
extreme events, and warming substantially larger than the assessed very likely range of future warming,
cannot be ruled out and are part of risk assessment.
{1.4, Cross-Chapter Box 1.3, 4.3, 4.4, 4.8, Cross-Chapter Box 4.1, 8.6, 9.2, Box 9.4, 11.8, Box 11.2, Cross-Chapter
Box 12.1} (Table SPM.1)
C.3.1 If global warming exceeds the assessed very likely range for a given GHG emissions scenario, including low GHG emissions
scenarios, global and regional changes in many aspects of the climate system, such as regional precipitation and other
CIDs, would also exceed their assessed very likely ranges (high confidence). Such low-likelihood, high-warming outcomes
are associated with potentially very large impacts, such as through more intense and more frequent heatwaves and heavy
precipitation, and high risks for human and ecological systems, particularly for high GHG emissions scenarios.
{Cross-Chapter Box 1.3, 4.3, 4.4, 4.8, Box 9.4, Box 11.2, Cross-Chapter Box 12.1, TS.1.4, Box TS.3, Box TS.4} (Table SPM.1)
C.3.2 Low-likelihood, high-impact outcomes34 could occur at global and regional scales even for global warming within the
very likely range for a given GHG emissions scenario. The probability of low-likelihood, high-impact outcomes increases SPM
with higher global warming levels (high confidence). Abrupt responses and tipping points of the climate system, such as
strongly increased Antarctic ice-sheet melt and forest dieback, cannot be ruled out (high confidence).
{1.4, 4.3, 4.4, 4.8, 5.4, 8.6, Box 9.4, Cross-Chapter Box 12.1, TS.1.4, TS.2.5, Box TS.3, Box TS.4, Box TS.9} (Table SPM.1)
C.3.3 If global warming increases, some compound extreme events18 with low likelihood in past and current climate will become
more frequent, and there will be a higher likelihood that events with increased intensities, durations and/or spatial extents
unprecedented in the observational record will occur (high confidence).
{11.8, Box 11.2, Cross-Chapter Box 12.1, Box TS.3, Box TS.9}
C.3.4 The Atlantic Meridional Overturning Circulation is very likely to weaken over the 21st century for all emissions scenarios.
While there is high confidence in the 21st century decline, there is only low confidence in the magnitude of the trend.
There is medium confidence that there will not be an abrupt collapse before 2100. If such a collapse were to occur, it
would very likely cause abrupt shifts in regional weather patterns and water cycle, such as a southward shift in the
tropical rain belt, weakening of the African and Asian monsoons and strengthening of Southern Hemisphere monsoons,
and drying in Europe.
{4.3, 8.6, 9.2, TS2.4, Box TS.3}
C.3.5 Unpredictable and rare natural events not related to human influence on climate may lead to low-likelihood, high-impact
outcomes. For example, a sequence of large explosive volcanic eruptions within decades has occurred in the past, causing
substantial global and regional climate perturbations over several decades. Such events cannot be ruled out in the future,
but due to their inherent unpredictability they are not included in the illustrative set of scenarios referred to in this Report
{2.2, Cross-Chapter Box 4.1, Box TS.3} (Box SPM.1)
D.1 From a physical science perspective, limiting human-induced global warming to a specific level requires
limiting cumulative CO2 emissions, reaching at least net zero CO2 emissions, along with strong reductions in
other greenhouse gas emissions. Strong, rapid and sustained reductions in CH4 emissions would also limit the
warming effect resulting from declining aerosol pollution and would improve air quality.
{3.3, 4.6, 5.1, 5.2, 5.4, 5.5, 5.6, Box 5.2, Cross-Chapter Box 5.1, 6.7, 7.6, 9.6} (Figure SPM.10, Table SPM.2)
27
Summary for Policymakers
D.1.1 This Report reaffirms with high confidence the AR5 finding that there is a near-linear relationship between cumulative
anthropogenic CO2 emissions and the global warming they cause. Each 1000 GtCO2 of cumulative CO2 emissions is assessed
to likely cause a 0.27°C to 0.63°C increase in global surface temperature with a best estimate of 0.45°C.41 This is a narrower
range compared to AR5 and SR1.5. This quantity is referred to as the transient climate response to cumulative CO2 emissions
(TCRE). This relationship implies that reaching net zero anthropogenic CO2 emissions42 is a requirement to stabilize
human-induced global temperature increase at any level, but that limiting global temperature increase to a specific level
would imply limiting cumulative CO2 emissions to within a carbon budget.43 {5.4, 5.5, TS.1.3, TS.3.3, Box TS.5} (Figure SPM.10)
SSP5-8.5
SPM The near-linear relationship
SSP3-7.0
2.5 between the cumulative
CO₂ emissions and global SSP2-4.5
warming for five illustrative
scenarios until year 2050 SSP1-2.6
2
SSP1-1.9
1.5
Historical global
warming
0.5
1950
2000
2019
2020
2030
2040
2050
HISTORICAL PROJECTIONS
Cumulative CO₂ emissions between 1850 and 2019 Cumulative CO₂ emissions between 2020 and 2050
Figure SPM.10 | Near-linear relationship between cumulative CO2 emissions and the increase in global surface temperature
Top panel: Historical data (thin black line) shows observed global surface temperature increase in °C since 1850–1900 as a function of historical cumulative carbon
dioxide (CO2) emissions in GtCO2 from 1850 to 2019. The grey range with its central line shows a corresponding estimate of the historical human-caused surface
warming (see Figure SPM.2). Coloured areas show the assessed very likely range of global surface temperature projections, and thick coloured central lines show the
median estimate as a function of cumulative CO2 emissions from 2020 until year 2050 for the set of illustrative scenarios (SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0, and
SSP5-8.5; see Figure SPM.4). Projections use the cumulative CO2 emissions of each respective scenario, and the projected global warming includes the contribution
from all anthropogenic forcers. The relationship is illustrated over the domain of cumulative CO2 emissions for which there is high confidence that the transient climate
response to cumulative CO2 emissions (TCRE) remains constant, and for the time period from 1850 to 2050 over which global CO2 emissions remain net positive under
all illustrative scenarios, as there is limited evidence supporting the quantitative application of TCRE to estimate temperature evolution under net negative CO2 emissions.
Bottom panel: Historical and projected cumulative CO2 emissions in GtCO2 for the respective scenarios.
{Section 5.5, Figure 5.31, Figure TS.18}
41 In the literature, units of °C per 1000 PgC (petagrams of carbon) are used, and the AR6 reports the TCRE likely range as 1.0°C to 2.3°C per 1000 PgC in the underlying report, with
a best estimate of 1.65°C.
42 The condition in which anthropogenic carbon dioxide (CO2) emissions are balanced by anthropogenic CO2 removals over a specified period (Glossary).
43 The term ‘carbon budget’ refers to the maximum amount of cumulative net global anthropogenic CO2 emissions that would result in limiting global warming to a given level with
a given probability, taking into account the effect of other anthropogenic climate forcers. This is referred to as the total carbon budget when expressed starting from the pre-industrial
period, and as the remaining carbon budget when expressed from a recent specified date (Glossary). Historical cumulative CO2 emissions determine to a large degree warming to
date, while future emissions cause future additional warming. The remaining carbon budget indicates how much CO2 could still be emitted while keeping warming below a specific
temperature level.
28
Summary for Policymakers
D.1.2 Over the period 1850–2019, a total of 2390 ± 240 (likely range) GtCO2 of anthropogenic CO2 was emitted. Remaining
carbon budgets have been estimated for several global temperature limits and various levels of probability, based on the
estimated value of TCRE and its uncertainty, estimates of historical warming, variations in projected warming from non-
CO2 emissions, climate system feedbacks such as emissions from thawing permafrost, and the global surface temperature
change after global anthropogenic CO2 emissions reach net zero.
{5.1, 5.5, Box 5.2, TS.3.3} (Table SPM.2)
Table SPM.2 | Estimates of historical carbon dioxide (CO2) emissions and remaining carbon budgets. Estimated remaining carbon budgets are
calculated from the beginning of 2020 and extend until global net zero CO2 emissions are reached. They refer to CO2 emissions, while accounting for the global
warming effect of non-CO2 emissions. Global warming in this table refers to human-induced global surface temperature increase, which excludes the impact
of natural variability on global temperatures in individual years.
{Table 3.1, 5.5.1, 5.5.2, Box 5.2, Table 5.1, Table 5.7, Table 5.8, Table TS.3}
a
Values at each 0.1°C increment of warming are available in Tables TS.3 and 5.8.
b
This likelihood is based on the uncertainty in transient climate response to cumulative CO2 emissions (TCRE) and additional Earth system feedbacks and provides the
probability that global warming will not exceed the temperature levels provided in the two left columns. Uncertainties related to historical warming (±550 GtCO2)
and non-CO2 forcing and response (±220 GtCO2) are partially addressed by the assessed uncertainty in TCRE, but uncertainties in recent emissions since 2015
(±20 GtCO2) and the climate response after net zero CO2 emissions are reached (±420 GtCO2) are separate.
c
Remaining carbon budget estimates consider the warming from non-CO2 drivers as implied by the scenarios assessed in SR1.5. The Working Group III Contribution
to AR6 will assess mitigation of non-CO2 emissions.
D.1.3 Several factors that determine estimates of the remaining carbon budget have been re-assessed, and updates to these
factors since SR1.5 are small. When adjusted for emissions since previous reports, estimates of remaining carbon budgets
are therefore of similar magnitude compared to SR1.5 but larger compared to AR5 due to methodological improvements.44
{5.5, Box 5.2, TS.3.3} (Table SPM.2)
D.1.4 Anthropogenic CO2 removal (CDR) has the potential to remove CO2 from the atmosphere and durably store it in reservoirs
(high confidence). CDR aims to compensate for residual emissions to reach net zero CO2 or net zero GHG emissions or, if
implemented at a scale where anthropogenic removals exceed anthropogenic emissions, to lower surface temperature.
CDR methods can have potentially wide-ranging effects on biogeochemical cycles and climate, which can either weaken
or strengthen the potential of these methods to remove CO2 and reduce warming, and can also influence water availability
and quality, food production and biodiversity45 (high confidence).
{5.6, Cross-Chapter Box 5.1, TS.3.3}
D.1.5 Anthropogenic CO2 removal (CDR) leading to global net negative emissions would lower the atmospheric CO2 concentration
and reverse surface ocean acidification (high confidence). Anthropogenic CO2 removals and emissions are partially
44 Compared to AR5, and when taking into account emissions since AR5, estimates in AR6 are about 300–350 GtCO2 larger for the remaining carbon budget consistent with limiting
warming to 1.5°C; for 2°C, the difference is about 400–500 GtCO2.
45 Potential negative and positive effects of CDR for biodiversity, water and food production are methods-specific and are often highly dependent on local context, management, prior
land use, and scale. IPCC Working Groups II and III assess the CDR potential and ecological and socio-economic effects of CDR methods in their AR6 contributions.
29
Summary for Policymakers
compensated by CO2 release and uptake respectively, from or to land and ocean carbon pools (very high confidence).
CDR would lower atmospheric CO2 by an amount approximately equal to the increase from an anthropogenic emission of
the same magnitude (high confidence). The atmospheric CO2 decrease from anthropogenic CO2 removals could be up to
10% less than the atmospheric CO2 increase from an equal amount of CO2 emissions, depending on the total amount of
CDR (medium confidence).
{5.3, 5.6, TS.3.3}
D.1.6 If global net negative CO2 emissions were to be achieved and be sustained, the global CO2-induced surface temperature
increase would be gradually reversed but other climate changes would continue in their current direction for decades to
millennia (high confidence). For instance, it would take several centuries to millennia for global mean sea level to reverse
course even under large net negative CO2 emissions (high confidence).
{4.6, 9.6, TS.3.3}
D.1.7 In the five illustrative scenarios, simultaneous changes in CH4, aerosol and ozone precursor emissions, which also
SPM contribute to air pollution, lead to a net global surface warming in the near and long term (high confidence). In the
long term, this net warming is lower in scenarios assuming air pollution controls combined with strong and sustained
CH4 emissions reductions (high confidence). In the low and very low GHG emissions scenarios, assumed reductions in
anthropogenic aerosol emissions lead to a net warming, while reductions in CH4 and other ozone precursor emissions
lead to a net cooling. Because of the short lifetime of both CH4 and aerosols, these climate effects partially counterbalance
each other, and reductions in CH4 emissions also contribute to improved air quality by reducing global surface ozone
(high confidence).
{6.7, Box TS.7} (Figure SPM.2, Box SPM.1)
D.1.8 Achieving global net zero CO2 emissions, with anthropogenic CO2 emissions balanced by anthropogenic removals of
CO2, is a requirement for stabilizing CO2-induced global surface temperature increase. This is different from achieving
net zero GHG emissions, where metric-weighted anthropogenic GHG emissions equal metric-weighted anthropogenic
GHG removals. For a given GHG emissions pathway, the pathways of individual GHGs determine the resulting climate
response,46 whereas the choice of emissions metric47 used to calculate aggregated emissions and removals of different
GHGs affects what point in time the aggregated GHGs are calculated to be net zero. Emissions pathways that reach and
sustain net zero GHG emissions defined by the 100-year global warming potential are projected to result in a decline in
surface temperature after an earlier peak (high confidence).
{4.6, 7.6, Box 7.3, TS.3.3}
D.2 Scenarios with very low or low GHG emissions (SSP1-1.9 and SSP1-2.6) lead within years to discernible effects
on greenhouse gas and aerosol concentrations and air quality, relative to high and very high GHG emissions
scenarios (SSP3-7.0 or SSP5-8.5). Under these contrasting scenarios, discernible differences in trends of global
surface temperature would begin to emerge from natural variability within around 20 years, and over longer
time periods for many other climatic impact-drivers (high confidence).
{4.6, 6.6, 6.7, Cross-Chapter Box 6.1, 9.6, 11.2, 11.4, 11.5, 11.6, Cross-Chapter Box 11.1, 12.4, 12.5} (Figure
SPM.8, Figure SPM.10)
D.2.1 Emissions reductions in 2020 associated with measures to reduce the spread of COVID-19 led to temporary but detectable
effects on air pollution (high confidence) and an associated small, temporary increase in total radiative forcing, primarily
due to reductions in cooling caused by aerosols arising from human activities (medium confidence). Global and regional
climate responses to this temporary forcing are, however, undetectable above natural variability (high confidence).
Atmospheric CO2 concentrations continued to rise in 2020, with no detectable decrease in the observed CO2 growth rate
(medium confidence).48
{Cross-Chapter Box 6.1, TS.3.3}
D.2.2 Reductions in GHG emissions also lead to air quality improvements. However, in the near term,49 even in scenarios with
strong reduction of GHGs, as in the low and very low GHG emissions scenarios (SSP1-2.6 and SSP1-1.9), these improvements
46 A general term for how the climate system responds to a radiative forcing (Glossary).
47 The choice of emissions metric depends on the purposes for which gases or forcing agents are being compared. This Report contains updated emissions metric values and assesses
new approaches to aggregating gases.
48 For other GHGs, there was insufficient literature available at the time of the assessment to assess detectable changes in their atmospheric growth rate during 2020.
49 Near term: 2021–2040.
30
Summary for Policymakers
are not sufficient in many polluted regions to achieve air quality guidelines specified by the World Health Organization
(high confidence). Scenarios with targeted reductions of air pollutant emissions lead to more rapid improvements in air
quality within years compared to reductions in GHG emissions only, but from 2040, further improvements are projected
in scenarios that combine efforts to reduce air pollutants as well as GHG emissions, with the magnitude of the benefit
varying between regions (high confidence).
{6.6, 6.7, Box TS.7}.
D.2.3 Scenarios with very low or low GHG emissions (SSP1-1.9 and SSP1-2.6) would have rapid and sustained effects to limit
human-caused climate change, compared with scenarios with high or very high GHG emissions (SSP3-7.0 or SSP5-8.5),
but early responses of the climate system can be masked by natural variability. For global surface temperature, differences
in 20-year trends would likely emerge during the near term under a very low GHG emissions scenario (SSP1-1.9), relative
to a high or very high GHG emissions scenario (SSP3-7.0 or SSP5-8.5). The response of many other climate variables would
emerge from natural variability at different times later in the 21st century (high confidence).
{4.6, Cross-Section Box TS.1} (Figure SPM.8, Figure SPM.10)
SPM
D.2.4 Scenarios with very low and low GHG emissions (SSP1-1.9 and SSP1-2.6) would lead to substantially smaller changes
in a range of CIDs36 beyond 2040 than under high and very high GHG emissions scenarios (SSP3-7.0 and SSP5-8.5).
By the end of the century, scenarios with very low and low GHG emissions would strongly limit the change of several
CIDs, such as the increases in the frequency of extreme sea level events, heavy precipitation and pluvial flooding, and
exceedance of dangerous heat thresholds, while limiting the number of regions where such exceedances occur, relative
to higher GHG emissions scenarios (high confidence). Changes would also be smaller in very low compared to low GHG
emissions scenarios, as well as for intermediate (SSP2-4.5) compared to high or very high GHG emissions scenarios (high
confidence).
{9.6, 11.2, 11.3, 11.4, 11.5, 11.6, 11.9, Cross-Chapter Box 11.1, 12.4, 12.5, TS.4.3}
31
Foreword
Technical
and Preface
Summary
33
34
TS Technical Summary
Coordinating Authors:
Paola A. Arias (Colombia), Nicolas Bellouin (United Kingdom/France), Erika Coppola (Italy),
Richard G. Jones (United Kingdom), Gerhard Krinner (France/Germany, France), Jochem Marotzke
(Germany), Vaishali Naik (United States of America), Matthew D. Palmer (United Kingdom),
Gian-Kasper Plattner (Switzerland), Joeri Rogelj (United Kingdom/Belgium), Maisa Rojas (Chile),
Jana Sillmann (Norway/Germany), Trude Storelvmo (Norway), Peter W. Thorne (Ireland/United
Kingdom), Blair Trewin (Australia)
Authors:
Krishna Achuta Rao (India), Bhupesh Adhikary (Nepal), Richard P. Allan (United Kingdom),
Kyle Armour (United States of America), Govindasamy Bala (India/United States of America),
Rondrotiana Barimalala (South Africa/Madagascar), Sophie Berger (France/Belgium),
Josep G. Canadell (Australia), Christophe Cassou (France), Annalisa Cherchi (Italy), William Collins
(United Kingdom), William D. Collins (United States of America), Sarah L. Connors (France/United
Kingdom), Susanna Corti (Italy), Faye Cruz (Philippines), Frank J. Dentener (EU/The Netherlands),
Claudine Dereczynski (Brazil), Alejandro Di Luca (Australia, Canada/Argentina), Aida Diongue
Niang (Senegal), Francisco J. Doblas-Reyes (Spain), Alessandro Dosio (Italy), Hervé Douville (France),
François Engelbrecht (South Africa), Veronika Eyring (Germany), Erich Fischer (Switzerland), Piers
Forster (United Kingdom), Baylor Fox-Kemper (United States of America), Jan S. Fuglestvedt
(Norway), John C. Fyfe (Canada), Nathan P. Gillett (Canada), Leah Goldfarb (France/United States
of America), Irina Gorodetskaya (Portugal/Russian Federation, Belgium), Jose Manuel Gutierrez
(Spain), Rafiq Hamdi (Belgium), Ed Hawkins (United Kingdom), Helene T. Hewitt (United Kingdom),
Pandora Hope (Australia), AKM Saiful Islam (Bangladesh), Christopher Jones (United Kingdom),
Darrell S. Kaufman (United States of America), Robert E. Kopp (United States of America),
Yu Kosaka (Japan), James Kossin (United States of America), Svitlana Krakovska (Ukraine),
June-Yi Lee (Republic of Korea), Jian Li (China), Thorsten Mauritsen (Sweden, Denmark), Thomas
K. Maycock (United States of America), Malte Meinshausen (Australia/Germany), Seung-Ki Min
(Republic of Korea), Pedro M. S. Monteiro (South Africa), Thanh Ngo-Duc (Vietnam), Friederike
Otto (United Kingdom/Germany), Izidine Pinto (South Africa/Mozambique), Anna Pirani (Italy),
Krishnan Raghavan (India), Roshanka Ranasinghe (The Netherlands/Sri Lanka, Australia), Alex
C. Ruane (United States of America), Lucas Ruiz (Argentina), Jean-Baptiste Sallée (France), Bjørn
H. Samset (Norway), Shubha Sathyendranath (UK/Canada, United Kingdom, Overseas Citizen of
India), Sonia I. Seneviratne (Switzerland), Anna A. Sörensson (Argentina), Sophie Szopa (France),
Izuru Takayabu (Japan), Anne-Marie Treguier (France), Bart van den Hurk (The Netherlands),
35
Technical Summary
Robert Vautard (France), Karina von Schuckmann (France/Germany), Sönke Zaehle (Germany),
Xuebin Zhang (Canada), Kirsten Zickfeld (Canada/Germany)
Contributing Authors:
Guðfinna Aðalgeirsdóttir (Iceland), Lincoln M. Alves (Brazil), Terje Berntsen (Norway),
Sara M. Blichner (Norway), Lisa Bock (Germany), Gregory G. Garner (United States of America),
Joelle Gergis (Australia), Sergey K. Gulev (Russian Federation), Mathias Hauser (Switzerland),
Flavio Lehner (United States of America/Switzerland), Chao Li (China), Marianne T. Lund
(Norway), Daniel J. Lunt (United Kingdom), Sebastian Milinski (Germany), Gemma Teresa
Narisma (Philippines), Zebedee R. J. Nicholls (Australia), Dirk Notz (Germany), Sophie Nowicki
(United States of America/France, United States of America), Bette Otto-Bliesner (United States
of America), Brodie Pearson (United States of America/United Kingdom), Adam S. Phillips
(United States of America), James Renwick (New Zealand), Stéphane Sénési (France), Lucas Silva
(Portugal/Switzerland), Aimee B. A. Slangen (The Netherlands), Thomas F. Stocker (Switzerland),
Claudia Tebaldi (United States of America), Laurent Terray (France), Sabin Thazhe Purayil (India),
Andrew Turner (United Kingdom), Steven Turnock (United Kingdom), Carolina Vera (Argentina),
Cunde Xiao (China), Panmao Zhai (China)
Review Editors:
Valérie Masson-Delmotte (France), Gregory M. Flato (Canada), Noureddine Yassa (Algeria)
TS
36
Technical Summary
Table of Contents
37
Technical Summary
1 In this Technical Summary, the following summary terms are used to describe the available evidence: limited, medium, or robust; and for the degree of agreement: low, medium, or high. A level of
confidence is expressed using five qualifiers: very low, low, medium, high, and very high, and typeset in italics, e.g., medium confidence. For a given evidence and agreement statement, different
confidence levels can be assigned, but increasing levels of evidence and degrees of agreement are correlated with increasing confidence (see Chapter 1, Box 1.1 for more details).
2 In this Technical Summary, the following terms are used to indicate the assessed likelihood of an outcome or a result: virtually certain 99–100% probability, very likely 90–100%, likely 66–100%,
about as likely as not 33–66%, unlikely 0–33%, very unlikely 0–10%, exceptionally unlikely 0–1%. Additional terms (extremely likely: 95–100%, more likely than not >50–100%, and extremely
unlikely 0–5%) may also be used when appropriate. Assessed likelihood is typeset in italics, e.g., very likely (see Chapter 1, Box 1.1 for more details). Throughout the WGI report and unless stated
otherwise, uncertainty is quantified using 90% uncertainty intervals. The 90% uncertainty interval, reported in square brackets [x to y], is estimated to have a 90% likelihood of covering the value
that is being estimated. The range encompasses the median value, and there is an estimated 10% combined likelihood of the value being below the lower end of the range (x) and above its upper
end (y). Often, the distribution will be considered symmetric about the corresponding best estimate, but this is not always the case. In this Report, an assessed 90% uncertainty interval is referred
to as a ‘very likely range’. Similarly, an assessed 66% uncertainty interval is referred to as a ‘likely range’.
3 The regional traceback matrices that provide the location of the assessment findings synthesized in Section TS.4 are in the Supplementary Material (SM) of Chapter 10.
4 Data archive is available at https://catalogue.ceda.ac.uk/uuid/3234e9111d4f4354af00c3aaecd879b7.
5 https://interactive-atlas.ipcc.ch/
6 The AR6 figures use one of the following approaches. For observations, the absence of ‘x’ symbols shows areas with statistical significance, while the presence of ‘x’ indicates non-significance.
For model projections, the method offers two approaches with varying complexity. In the simple approach, high agreement (≥80%) is indicated with no overlay, and diagonal lines (///) show low
agreement (<80%); In the advanced approach, areas with no overlay display robust signal (≥66% of models show change greater than the variability threshold and ≥80% of all models agree on
the sign of change), reverse diagonal lines (\\\) show no robust signal, and crossed lines show conflicting signals (i.e., significant change but low agreement). Cross-Chapter Box Atlas.1 provides
more information on the AR6 method for visualizing robustness and uncertainty on maps.
38
Technical Summary
This box provides short descriptions of key concepts that are relevant to the AR6 WGI assessment, with a focus on their use in the
Technical Summary and the Summary for Policymakers. The Glossary (Annex VII) includes more information on these concepts along
with definitions of many other important terms and concepts used in this Report.
Global warming: Global warming refers to the change of global surface temperature relative to a baseline depending upon the
application. Specific global warming levels, such as 1.5°C, 2°C, 3°C or 4°C, are defined as changes in global surface temperature
relative to the years 1850–1900 as the baseline (the earliest period of reliable observations with sufficient geographic coverage). They
are used to assess and communicate information about global and regional changes, linking to scenarios and used as a common basis
for Working Group II (WGII) and Working Group III (WGIII) assessments. (Section TS.1.3, Cross-Section Box TS.1) {1.4.1, 1.6.2, 4.6.1,
Cross-Chapter Boxes 1.5, 2.3, 11.1, and 12.1, Atlas Sections 3–11, Glossary}
Emergence: Emergence refers to the experience or appearance of novel conditions of a particular climate variable in a given region.
This concept is often expressed as the ratio of the change in a climate variable relative to the amplitude of natural variations of that
variable (often termed a ‘signal-to-noise’ ratio, with emergence occurring at a defined threshold of this ratio). Emergence can be
expressed in terms of a time or a global warming level at which the novel conditions appear and can be estimated using observations
or model simulations. (Sections TS.1.2.3 and TS.4.2) {1.4.2, FAQ 1.2, 7.5.5, 10.3, 10.4, 12.5.2, Cross-Chapter Box Atlas.1, Glossary}
Cumulative carbon dioxide (CO2) emissions: The total net amount of CO2 emitted into the atmosphere as a result of human TS
activities. Given the nearly linear relationship between cumulative CO2 emissions and increases in global surface temperature,
cumulative CO2 emissions are relevant for understanding how past and future CO2 emissions affect global surface temperature.
A related term – remaining carbon budget – is used to describe the total net amount of CO2 that could be released in the future by
human activities while keeping global warming to a specific global warming level, such as 1.5°C, taking into account the warming
contribution from non-CO2 forcers as well. The remaining carbon budget is expressed from a recent specified date, while the total
carbon budget is expressed starting from the pre-industrial period. (Sections TS.1.3 and TS.3.3) {1.6.3, 5.5, Glossary}
Net zero CO2 emissions: A condition that occurs when the amount of CO2 emitted into the atmosphere by human activities equals
the amount of CO2 removed from the atmosphere by human activities over a specified period of time. Net negative CO2 emissions
occur when anthropogenic removals exceed anthropogenic emissions. (Section TS.3.3) {Box 1.4, Glossary}
Earth’s energy imbalance: In a stable climate, the amount of energy that Earth receives from the Sun is approximately in balance
with the amount of energy that is lost to space in the form of reflected sunlight and thermal radiation. ‘Climate drivers’, such as an
increase in greenhouse gases or aerosols, interfere with this balance, causing the system to either gain or lose energy. The strength
of a climate driver is quantified by its effective radiative forcing (ERF), measured in W m–2. Positive ERF leads to warming, and
negative ERF leads to cooling. That warming or cooling in turn can change the energy imbalance through many positive (amplifying)
or negative (dampening) climate feedbacks. (Sections TS.2.2, TS.3.1 and TS.3.2) {2.2.8, 7.2, 7.3, 7.4, Box 7.1, Box 7.2, Glossary}
Attribution: Attribution is the process of evaluating the relative contributions of multiple causal factors to an observed change in
climate variables (e.g., global surface temperature, global mean sea level), or to the occurrence of extreme weather or climate-related
events. Attributed causal factors include human activities (such as increases in greenhouse gas concentration and aerosols, or land-use
change) or natural external drivers (solar and volcanic influences), and in some cases internal variability. (Sections TS.1.2.4 and TS.2,
Box TS.10) {Cross-Working Group Box: Attribution in Chapter 1; 3.5; 3.8; 10.4; 11.2.4; Glossary}
Committed change, long-term commitment: Changes in the climate system, resulting from past, present and future human
activities, which will continue long into the future (centuries to millennia) even with strong reductions in greenhouse gas emissions.
Some aspects of the climate system, including the terrestrial biosphere, the deep ocean and the cryosphere, respond much more slowly
than surface temperatures to changes in greenhouse gas concentrations. As a result, there are already substantial committed changes
associated with past greenhouse gas emissions. For example, global mean sea level will continue to rise for thousands of years, even
if future CO2 emissions are reduced to net zero and global warming halted, as excess energy due to past emissions continues to
propagate into the deep ocean and as glaciers and ice sheets continue to melt. (Section TS.2.1, Box TS.4, Box TS.9) {1.2.1, 1.3, Box 1.2,
Cross-Chapter Box 5.3}
39
Technical Summary
Distillation: The process of synthesizing information about climate change from multiple lines of evidence obtained from a variety
of sources, taking into account user context and values. It leads to an increase in the usability, usefulness and relevance of climate
information, enhances stakeholder trust, and expands the foundation of evidence used in climate services. It is particularly relevant in
the context of co-producing regional-scale climate information to support decision-making. (Section TS.4.1, Box TS.11) {10.1, 10.5, 12.6}
(Climate change) risk: The concept of risk is a key aspect of how the IPCC assesses and communicates to decision-makers about the
potential for adverse consequences for human or ecological systems, recognizing the diversity of values and objectives associated with
such systems. In the context of climate change, risks can arise from potential impacts of climate change as well as human responses to
climate change. WGI contributes to the common IPCC risk framing through the assessment of relevant climate information, including
climatic impact-drivers and low-likelihood, high-impact outcomes. (Sections TS.1.4 and TS.4.1, Box TS.4) {Cross-Chapter Boxes 1.3 and
12.1, Glossary}
Climatic impact-drivers: Physical climate system conditions (e.g., means, events, extremes) that can be directly connected with
having impacts on human or ecological systems are described as ‘climatic impact-drivers’ (CIDs) without anticipating whether their
impacts are detrimental (i.e., as for hazards in the context of climate change risks) or provide potential opportunities. A range of indices
may capture the sector- or application-relevant characteristics of a climatic impact-driver and can reflect exceedances of identified
tolerance thresholds. (Sections TS.1.4 and TS.4.3) {12.1–12.3, FAQ 12.1, Glossary}
TS
Storylines: The term storyline is used both in connection to scenarios (related to a future trajectory of emissions or socio-economic
developments) or to describe plausible trajectories of weather and climate conditions or events, especially those related to high levels
of risk. Physical climate storylines are introduced in AR6 to explore uncertainties in climate change and natural climate variability, to
develop and communicate integrated and context-relevant regional climate information, and to address issues with deep uncertainty7,
including low-likelihood, high-impact outcomes. (Section TS.1.4, Box TS.3, Infographic TS.1) {1.4.4, Box 10.2, Glossary}
Low-likelihood, high impact outcomes: Outcomes/events whose probability of occurrence is low or not well known (as in the context
of deep uncertainty) but whose potential impacts on society and ecosystems could be high. To better inform risk assessment and
decision-making, such low-likelihood outcomes are considered if they are associated with very large consequences and may therefore
constitute material risks, even though those consequences do not necessarily represent the most likely outcome. (Section TS.1.4,
Box TS.3, Figure TS.6) {1.4.4, 4.8, Cross Chapter Box 1.3, Glossary}
As part of the AR6 cycle, the IPCC produced three Special Reports The Report has been peer-reviewed by the scientific community
in 2018 and 2019: the Special Report on Global Warming of and governments (Annex X provides the Expert Reviewer list). The
1.5°C (SR1.5), the Special Report on the Ocean and Cryosphere in substantive introduction provided by Chapter 1 is followed by a first
a Changing Climate (SROCC), and the Special Report on Climate set of chapters dedicated to large-scale climate knowledge (Chapters
Change and Land (SRCCL). 2–4), which encompasses observations and paleoclimate evidence,
causes of observed changes, and projections; these are complemented
The AR6 WGI Report provides a full and comprehensive assessment by Chapter 11 for large-scale changes in extremes. The second set of
of the physical science basis of climate change that builds on the chapters (Chapters 5–9) is orientated around the understanding of
previous assessments and these Special Reports and considers new key climate system components and processes, including the global
information and knowledge from the recent scientific literature8, cycles of carbon, energy and water; short-lived climate forcers and their
including longer observational datasets and new scenarios and link to air quality; and the ocean, cryosphere and sea level change.
model results. The last set of chapters (Chapters 10–12 and the Atlas) is dedicated to
the assessment and distillation of regional climate information from
The structure of the AR6 WGI Report is designed to enhance the multiple lines of evidence at sub-continental to local scales (including
visibility of knowledge developments and to facilitate the integration urban climate), with a focus on recent and projected regional changes
of multiple lines of evidence, thereby improving confidence in findings. in mean climate, extremes, and climatic impact-drivers. The new online
7 Although not a core concept of the WGI Report, deep uncertainty is used in the Technical Summary in the following sense: ‘A situation of deep uncertainty exists when experts or stakeholders do
not know or cannot agree on: (1) appropriate conceptual models that describe relationships among key driving forces in a system; (2) the probability distributions used to represent uncertainty
about key variables and parameters; and/or (3) how to weigh and value desirable alternative outcomes’ (Lempert et al., 2003). Lempert, R. J., Popper, S. W., and Bankes, S. C. (2003). Shaping the
next one hundred years: New methods for quantitative long-term strategy analysis (MR-1626-RPC). Santa Monica, CA: The RAND Pardee Center.
8 The assessment covers scientific literature accepted for publication by 31 January 2021.
40
Technical Summary
Interactive Atlas allows users to interact in a flexible manner through • Updated assessment of recent warming: The AR5 reported
maps, time series and summary statistics with climate information for a smaller rate of increase in global mean surface temperature over
a set of updated WGI reference regions. The Report also includes 34 the period 1998–2012 than the rate calculated since 1951. Based
Frequently Asked Questions and answers for the general public (https:// on updated observational datasets showing a larger trend over
www.ipcc.ch/report/ar6/wg1/faqs). 1998–2012 than earlier estimates, there is now high confidence
that the observed 1998–2012 global surface temperature trend
Together, this Technical Summary and the underlying chapters aim at is consistent with ensembles of climate model simulations,
providing a comprehensive picture of knowledge progress since the and there is now very high confidence that the slower rate of
WGI contribution to AR5. Multiple lines of scientific evidence confirm global surface temperature increase observed over this period
that the climate is changing due to human influence. Important was a temporary event induced by internal and naturally forced
advances in the ability to understand past, present and possible variability that partly offset the anthropogenic surface warming
future changes should result in better-informed decision-making. trend over this period, while heat uptake continued to increase
in the ocean. Since 2012, strong warming has been observed,
Some of the new results and main updates to key findings in this with the past five years (2016–2020) being the hottest five-
Report compared to AR5, SR1.5, SRCCL, and SROCC are summarized year period in the instrumental record since at least 1850 (high
below. Relevant Technical Summary sections with further details are confidence). (Section TS.1.2, Cross-Section Box TS.1)
shown in parentheses after each bullet point. • Magnitude of climate system response: In this Report, it has
been possible to reduce the long-standing uncertainty ranges
for metrics that quantify the response of the climate system to
Selected Updates and/or New Results since AR5 radiative forcing, such as the equilibrium climate sensitivity (ECS)
and the transient climate response (TCR), due to substantial
• Human influence9 on the climate system is now an advances (e.g., a 50% reduction in the uncertainty range of cloud
established fact: The Fourth Assessment Report (AR4) stated feedbacks) and improved integration of multiple lines of evidence, TS
in 2007 that ‘warming of the climate system is unequivocal’, and including paleoclimate information. Improved quantification
AR5 stated in 2013 that ‘human influence on the climate system of ERF, the climate system radiative response, and the observed
is clear’. Combined evidence from across the climate system energy increase in the Earth system over the past five decades
strengthens this finding. It is unequivocal that the increase of demonstrate improved consistency between independent
CO2, methane (CH4) and nitrous oxide (N2O) in the atmosphere estimates of climate drivers, the combined climate feedbacks, and
over the industrial era is the result of human activities and that the observed energy increase relative to AR5. (Section TS.3.2)
human influence is the main driver10 of many changes observed • Improved constraints on projections of future climate
across the atmosphere, ocean, cryosphere and biosphere. change: For the first time in an IPCC report, the assessed future
(Sections TS.1.2, TS.2.1 and TS.3.1) change in global surface temperature is consistently constructed
• Observed global warming to date: A combination of by combining scenario-based projections (which AR5 focused
improved observational records and a series of very warm years on) with observational constraints based on past simulations
since AR5 have resulted in a substantial increase in the estimated of warming as well as the updated assessment of ECS and TCR.
level of global warming to date. The contribution of changes in In addition, initialized forecasts have been used for the period
observational understanding alone between AR5 and AR6 leads 2019–2028. The inclusion of these lines of evidence reduces
to an increase of about 0.1°C in the estimated warming since the assessed uncertainty for each scenario. (Section TS.1.3,
1850–1900. For the decade 2011–2020, the increase in global Cross-Section Box TS.1)
surface temperature since 1850–1900 is assessed to be 1.09 • Air quality: The AR5 assessed that projections of air quality
[0.95 to 1.20] °C.11 Estimates of crossing times of global warming are driven primarily by precursor emissions, including CH4. New
levels and estimates of remaining carbon budgets are updated scenarios explore a diversity of future options in air pollution
accordingly. (Section TS.1.2, Cross-Section Box TS.1) management. The AR6 reports rapid recent shifts in the
• Paleoclimate evidence: The AR5 assessed that many of the geographical distribution of some of these precursor emissions,
changes observed since the 1950s are unprecedented over confirms the AR5 finding, and shows higher warming effects
decades to millennia. Updated paleoclimate evidence strengthens of short-lived climate forcers in scenarios with the highest air
this assessment; over the past several decades, key indicators of pollution. (Sections TS.1.3 and TS.2.2, Box TS.7)
the climate system are increasingly at levels unseen in centuries • Effects of short-lived climate forcers on global warming:
to millennia and are changing at rates unprecedented in at least The AR5 assessed the radiative forcing for emitted compounds.
the last 2000 years. (Box TS.2, Section TS.2) The AR6 has extended this by assessing the emissions-based ERFs
9 Human influence on the climate system refers to human-driven activities that lead to changes in the climate system due to perturbations of Earth’s energy budget (also called anthropogenic
forcing). Human influence results from emissions of greenhouse gases, aerosols and tropospheric ozone precursors, ozone-depleting substances, and land-use change.
10 Throughout this Technical Summary, ‘main driver’ means responsible for more than 50% of the change.
11 Throughout the WGI report and unless stated otherwise, uncertainty is quantified using 90% uncertainty intervals. The 90% uncertainty interval, reported in square brackets [x to y], is estimated
to have a 90% likelihood of covering the value that is being estimated. The range encompasses the median value and there is an estimated 10% combined likelihood of the value being below the
lower end of the range (x) and above its upper end (y). Often the distribution will be considered symmetric about the corresponding best estimate, but this is not always the case. In this Report,
an assessed 90% uncertainty interval is referred to as a ‘very likely range’. Similarly, an assessed 66% uncertainty interval is referred to as a ‘likely range’.
41
Technical Summary
also accounting for aerosol–cloud interactions. The best estimates • Effect of short-lived climate forcers on global warming
of ERF attributed to sulphur dioxide (SO2) and CH4 emissions are in coming decades: The SR1.5 stated that reductions in
substantially greater than in AR5, while that of black carbon is emissions of cooling aerosols partially offset greenhouse gas
substantially reduced. The magnitude of uncertainty in the ERF mitigation effects for two to three decades in pathways limiting
due to black carbon emissions has also been reduced relative to global warming to 1.5°C. The AR6 assessment updates the AR5
AR5. (Section TS.3.1) assessment of the net cooling effect of aerosols and confirms
• Global water cycle: The AR5 assessed that anthropogenic that changes in short-lived climate forcers will very likely cause
influences have likely affected the global water cycle since further warming in the next two decades across all scenarios.
1960. The dedicated chapter in AR6 (Chapter 8) concludes with (Section TS.1.3, Box TS.7)
high confidence that human-caused climate change has driven • COVID-19: Temporary emissions reductions in 2020 associated
detectable changes in the global water cycle since the mid-20th with COVID-19 containment led to small and positive net radiative
century, with a better understanding of the response to aerosol effect (warming influence). However, global and regional climate
and greenhouse gas changes. The AR6 further projects with high responses to this forcing are undetectable above internal climate
confidence an increase in the variability of the water cycle in most variability due to the temporary nature of emissions reductions.
regions of the world and under all emissions scenarios. (Box TS.6) (Section TS.3.3)
• Extreme events: The AR5 assessed that human influence
had been detected in changes in some climate extremes.
A dedicated chapter in AR6 (Chapter 11) concludes that it is Selected Updates and/or New Results Since AR5, SRCCL
now an established fact that human-induced greenhouse gas and SROCC
emissions have led to an increased frequency and/or intensity
of some weather and climate extremes since 1850, in particular • Atmospheric concentration of methane: The SRCCL reported
for temperature extremes. Evidence of observed changes and a resumption of atmospheric CH4 concentration growth since
TS attribution to human influence has strengthened for several types 2007. The AR6 reports a faster growth over 2014–2019 and
of extremes since AR5, in particular for extreme precipitation, assesses growth since 2007 to be largely driven by emissions
droughts, tropical cyclones and compound extremes (including from the fossil fuels and agriculture (dominated by livestock)
fire weather). (Sections TS.1.2 and TS.2.1, Box TS.10) sectors. (Section TS.2.2)
• Land and ocean carbon sinks: The SRCCL assessed that the
persistence of the land carbon sink is uncertain due to climate
Selected Updates and/or New Results Since AR5 and SR1.5 change. The AR6 finds that land and ocean carbon sinks are
projected to continue to grow until 2100 with increasing
• Timing of crossing 1.5°C global warming: Slightly different atmospheric concentrations of CO2, but the fraction of emissions
approaches are used in SR1.5 and in this Report. SR1.5 assessed taken up by land and ocean is expected to decline as the CO2
a likely range of 2030 to 2052 for reaching a global warming concentration increases, with a much larger uncertainty range for
level of 1.5°C (for a 30-year period), assuming a continued, the land sink. The AR5, SR1.5 and SRCCL assessed carbon dioxide
constant rate of warming. In AR6, combining the larger estimate removal options and scenarios. The AR6 finds that the carbon
of global warming to date and the assessed climate response to cycle response is asymmetric for pulse emissions or removals,
all considered scenarios, the central estimate of crossing 1.5°C of which means that CO2 emissions would be more effective at
global warming (for a 20-year period) occurs in the early 2030s, in raising atmospheric CO2 than CO2 removals are at lowering
the early part of the likely range assessed in SR1.5, assuming no atmospheric CO2. (Section TS.3.3, Box TS.5)
major volcanic eruption. (Section TS.1.3, Cross-Section Box TS.1) • Ocean stratification increase12 : Refined analyses of available
• Remaining carbon budgets: The AR5 had assessed the observations in the AR6 lead to a reassessment of the rate of
transient climate response to cumulative emissions of CO2 to be increase of the global stratification in the upper 200 m to be double
likely in the range of 0.8°C to 2.5°C per 1000 GtC (1 Gigatonne that estimated in SROCC from 1970 to 2018. (Section TS.2.4)
of carbon, GtC, = 1 Petagram of carbon, PgC, = 3.664 Gigatonnes • Projected ocean oxygen loss: Future subsurface oxygen
of carbon dioxide, GtCO2), and this was also used in SR1.5. The decline in new projections assessed in WGI AR6 is substantially
assessment in AR6, based on multiple lines of evidence, leads to greater in 2080–2099 than assessed in SROCC. (Section TS.2.4)
a narrower likely range of 1.0°C–2.3°C per 1000 GtC. This has • Ice loss from glaciers and ice sheets: Since SROCC, globally
been incorporated in updated estimates of remaining carbon resolved glacier changes have improved estimates of glacier mass
budgets (see Section TS.3.3.1), together with methodological loss over the past 20 years, and estimates of the Greenland and
improvements and recent observations. (Sections TS.1.3 Antarctic Ice Sheet loss have been extended to 2020. (Section TS.2.5)
and TS.3.3) • Observed global mean sea level change: new observation-
based estimates published since SROCC lead to an assessed sea
level rise estimate from 1901 to 2018 that is now consistent with
the sum of individual components and consistent with closure of
the global energy budget. (Box TS.4)
12 Increased stratification reduces the vertical exchange of heat, salinity, oxygen, carbon and nutrients. Stratification is an important indicator for ocean circulation.
42
Technical Summary
• Projected global mean sea level change: The AR6 projections TS.1 A Changing Climate
of global mean sea level are based on projections from ocean
thermal expansion and land ice contribution estimates, which are This section introduces the assessment of the physical science basis of
consistent with the assessed ECS and assessed changes in global climate change in the AR6 and presents the climate context in which
surface temperature. They are underpinned by new land ice model this assessment takes place, recent progress in climate science and
intercomparisons and consideration of processes associated with the relevance of global and regional climate information for impact
low confidence to characterize the deep uncertainty in future ice and risk assessments. The future emissions scenarios and global
loss from Antarctica. The AR6 projections based on new models warming levels, used to integrate assessments across this Report,
and methods are broadly consistent with SROCC findings. are introduced and their applications for future climate projections
(Box TS.4) are briefly addressed. Paleoclimate science provides a long-term
context for observed climate change of the past 150 years and the
projected changes in the 21st century and beyond (Box TS.2). The
assessment of past, current and future global surface temperature
changes relative to the standard baselines and reference periods13
used throughout this Report is summarized in Cross-Section Box TS.1.
Earth’s climate system has evolved over many millions of years, and
evidence from natural archives provides a long-term perspective on
observed changes and projected changes over the coming centuries.
These reconstructions of past climate also show that atmospheric
CO2 concentrations and global surface temperature are strongly
coupled (Figure TS.1), based on evidence from a variety of proxy
records over multiple time scales (Box TS.2, Section TS.2). Levels of
global warming (see Core Concepts Box) that have not been seen
in millions of years could be reached by 2300, depending on the
emissions pathway that is followed (Section TS.1.3). For example,
there is medium confidence that, by 2300, an intermediate scenario14
used in this Report leads to global surface temperatures of [2.3°C
to 4.6°C] higher than 1850–1900, similar to the mid-Pliocene Warm
13 Several baselines or reference periods are used consistently throughout this Report. Baseline refers to a period against which anomalies (i.e., differences from the average value for the baseline
period) are calculated. Examples include the 1750 baseline (used for anthropogenic radiative forcings), the 1850–1900 baseline (an approximation for pre-industrial global surface temperature
from which global warming levels are calculated) and the 1995–2014 baseline (used for many climate model projections). A reference period indicates a time period over which various statistics
are calculated (e.g., the near-term reference period, 2021–2040). Paleo reference periods are listed in Box TS.2. {1.4.1, Cross-Chapter Boxes 1.2 and 2.1}
14 Please refer to Section TS.1.3.1 for an overview of the climate change scenarios used in this Report.
43
Technical Summary
Atmospheric CO2 concentration (ppm) Atmospheric CO2 concentration and global surface temperature change
during the last 60 million years and projections for the next 300 years SSP5-8.5
2000 SSP5-8.5
1000
SSP2-4.5 2300
400
SSP1-2.6
200
2100
early Eocene mid-Pliocene 2020
SSP1-2.6
Global surface temperature change (°C)
15
SSP5-8.5 2300
10
(relative to 1850–1900)
5 SSP2-4.5
0 SSP1-2.6 2100
TS
-5
0 2 4 6 8 10 12 14 16 18
60 50 40 30 20 10 9 7 5 3 1 800 600 400 200 0 1850 2000 2150 2300
Temperature (°C)
Millions of years Thousands of years Year CE relative to 1850–1900
Figure TS.1 | Changes in atmospheric CO2 and global surface temperature (relative to 1850–1900) from the deep past to the next 300 years. The intent of
this figure is to show that CO2 and temperature covary, both in the past and into the future, and that projected CO2 and temperatures are similar to those only from many millions
of years ago. CO2 concentrations from millions of years ago are reconstructed from multiple proxy records (grey dots are data from Section 2.2.3.1, Figure 2.3 shown with cubic-
spline fit). CO2 levels for the last 800,000 years through the mid-20th century are from air trapped in polar ice; recent values are from direct air measurements. Global surface
temperature prior to 1850 is estimated from marine oxygen isotopes, one of multiple sources of evidence used to assess paleo temperatures in this Report. Temperature of the
past 170 years is the AR6 assessed mean. CO2 levels and global surface temperature change for the future are shown for three Shared Socio-economic Pathway (SSP) scenarios
through 2300 CE, using Earth system model emulators calibrated to the assessed global surface temperatures. Their smooth trajectories do not account for inter-annual to
inter-decadal variability, including transient response to potential volcanic eruptions. Global maps for two paleo reference periods are based on Coupled Model Intercomparison
Project Phase 6 (CMIP6) and pre-CMIP6 multi-model means, with site-level proxy data for comparison (squares and circles are marine and terrestrial, respectively). The map
for 2020 is an estimate of the total observed warming since 1850–1900. Global maps at right show two SSP scenarios at 2100 (2081–2100) and at 2300 (2281–2300; map
from CMIP6 models; temperature assessed in 4.7.1). A brief account of the major climate forcings associated with past global temperature changes is in Cross-Chapter Box 2.1.
(Section TS.1.3, Figure TS.9, Cross-Section Box TS.1, Box TS.2) {1.2.1.2; Figures 1.14 and 1.5; 2.2.3; 2.3.1.1; 2.3.1.1.1; Figures 2.4 and 2.5; Cross-Chapter Box 2.1, Figure 1;
4.5.1; 4.7.1; Cross-Chapter Box 4.1; Cross-Chapter Box 7.1; Figure 7.13}
Period [2.5°C to 4°C], about 3.2 million years ago, whereas the high been studied systematically since the early 20th century. Other
CO2 emissions scenario SSP5-8.5 leads to temperatures of [6.6°C major anthropogenic drivers, such as atmospheric aerosols (fine
to 14.1°C] by 2300, which overlaps with the Early Eocene Climate solid particles or liquid droplets), land-use change and non-CO2
Optimum [10°C to 18°C], about 50 million years ago. {Cross-Chapter greenhouse gases, were identified by the 1970s. Since systematic
Boxes 2.1 and 2.4, 2.3.1, 4.3.1.1, 4.7.1.2, 7.4.4.1} scientific assessments began in the 1970s, the influence of human
activities on the warming of the climate system has evolved from
Understanding of the climate system’s fundamental elements is theory to established fact (see also Section TS.2). The evidence for
robust and well established. Scientists in the 19th century identified human influence on recent climate change strengthened from the
the major natural factors influencing the climate system. They IPCC First Assessment Report in 1990 to the IPCC Fifth Assessment
also hypothesized the potential for anthropogenic climate change Report in 2013/14, and is now even stronger in this assessment
due to CO2 emitted by combustion of fossil fuels (petroleum, (Sections TS.1.2.4 and TS.2). Changes across a greater number of
coal, natural gas). The principal natural drivers of climate change, climate system components, including changes in regional climate
including changes in incoming solar radiation, volcanic activity, and extremes can now be attributed to human influence (see
orbital cycles and changes in global biogeochemical cycles, have Sections TS.2 and TS.4). {1.3.1–1.3.5, 3.1, 11.2, 11.9}
44
Technical Summary
Paleoclimate evidence is integrated within multiple lines of evidence across the WGI Report to more fully understand
the climate system. Paleo evidence extends instrument-based observations of climate variables and climate drivers
back in time, providing the long-term context needed to gauge the extent to which recent and potential future
changes are unusual (Section TS.2, Figure TS.1). Pre-industrial climate states complement evidence from climate model
projections by providing real-world examples of climate characteristics for past global warming levels, with empirical
evidence for how the slow-responding components of the climate system operate over centuries to millennia – the
time scale for committed climate change (Core Concepts Box, Box TS.4, Box TS.9). Information about the state of
the climate system during well-described paleoclimate reference periods helps narrow the uncertainty range in the
overall assessment of Earth’s sensitivity to climate forcing (Section TS.3.2.1). {Cross-Chapter Box 2.1, FAQ 1.3, FAQ 2.1}
Paleoclimate reference periods. Over the long evolution of Earth’s climate, several periods have received extensive research
attention as examples of distinct climate states and rapid climate transitions (Box TS.2, Figure 1). These paleoclimate reference periods
represent the present geological era (Cenozoic; past 65 million years) and are used across chapters to help structure the assessment
of climate changes prior to industrialization. Cross-Chapter Box 2.1 describes the reference periods, along with a brief account of their
climate forcings, and lists where each is discussed in other chapters. Cross-Chapter Box 2.4 summarizes information on one of the
reference periods, the mid-Pliocene Warm Period. The Interactive Atlas includes model output from the World Climate Research
Programme Coupled Model Intercomparison Project Phase 6 (CMIP6) for four of the paleoclimate reference periods.
Three selected global climate indicators covary across multiple paleoclimate reference periods
TS
(a) (b)
Reference period CO2 Temperature Sea level
(*See Interactive Atlas for climate model output) Age (ppm) (°C) (m)
2000
Recent past 1995–2014 CE 360 397 0.66 to 1.00 0.15 to 0.25 Early Eocene
Atmospheric CO2 (ppm)
Approximate pre-industrial 1850–1900 CE 286 296 –0.15 to +0.11 –0.03 to 0.00
1000
Last Millennium 850–1850 CE 278 to 285 –0.14 ~ 0.24 –0.05 to 0.03
Mid-Holocene* 6.5–5.5 ka 260 to 268 0.2 to 1.0 –3.5 to +0.5
Last Deglacial Transition 500
18–11 ka 193 271 not assessed –120 –50 Recent past Mid-Pliocene
Last Glacial Maximum* 23–19 ka 188 to 194 –5 to –7 –134 to –125 1850–1900
Last Interglacial
Last Interglacial* 129–116 ka 266 to 282 0.5 to 1.5 5 to 10
200 Mid-Holocene
Mid-Pliocene Warm Period* 3.3–3.0 Ma 360 to 420 2.5 to 4.0 5 to 25
Last Glacial Maximum
Early Eocene 53–49 Ma 1150 to 2500 10 to 18 70 to 76
Paleocene-Eocene Thermal Maximum 55.9–55.7 Ma 900 2000 10 to 25 not assessed 100
–10 –5 0 5 10 15 20
X to Y: very likely range (caveats in Figure 2.34) lower 1850–1900 higher
X Y: start to end of period, with no stated uncertainty Global surface temperature (°C)
relative to 1850–1900
X ~ Y: lowest and highest values, with not stated uncertainty
Box TS.2, Figure 1 | Paleoclimate and recent reference periods, with selected key indicators. The intent of this figure is to list the paleoclimate reference
periods used in this Report, to summarize three key global climate indicators, and compare CO2 with global temperature over multiple periods. (a) Three large-scale
climate indicators (atmospheric CO2, global surface temperature relative to 1850–1900, and global mean sea level relative to 1900), based on assessments in
Chapter 2, with confidence levels ranging from low to very high. (b) Comparison between global surface temperature (relative to 1850–1900) and atmospheric CO2
concentration (shown on a log scale) for multiple reference periods (mid-points with 5–95% ranges). {2.2.3, 2.3.1.1, 2.3.3.3, Figure 2.34}
Paleoclimate models and reconstructions. Climate models that target paleoclimate reference periods have been featured by
the IPCC since the First Assessment Report. Under the framework of CMIP6-PMIP4 (Paleoclimate Modelling Intercomparison Project),
new protocols for model intercomparisons have been developed for multiple paleoclimate reference periods. These modelling efforts
have led to improved understanding of the climate response to different external forcings, including changes in Earth’s orbital and
plate movements, solar irradiance, volcanism, ice-sheet size and atmospheric greenhouse gases. Likewise, quantitative reconstructions
of climate variables from proxy records that are compared with paleoclimate simulations have improved as the number of study sites
and variety of proxy types have expanded, and as records have been compiled into new regional and global datasets. {1.3.2, 1.5.1,
Cross-Chapter Boxes 2.1 and 2.4}
Global surface temperature. Since AR5, updated climate forcings, improved models, new understanding of the strengths
and weaknesses of a growing array of proxy records, better chronologies and more robust proxy data products have led to better
agreement between models and reconstructions. For global surface temperature, the mid-point of the AR6-assessed range and the
median of the model-simulated temperatures differ by an average of 0.5°C across five reference periods; they overlap within their
45
Technical Summary
90% ranges in four of five cases, which together span from about 6 [5 to 7]°C colder during the Last Glacial Maximum to about 14 [10
to 18] °C warmer during the Early Eocene, relative to 1850–1900 (Box TS.2, Figure 2a). Changes in temperature by latitude in response
to multiple forcings show that polar amplification (stronger warming at high latitudes than the global average) is a prominent feature
of the climate system across multiple climate states, and the ability of models to simulate this polar amplification in past warm
climates has improved since AR5 (high confidence). Over the past millennium, and especially since about 1300 CE, simulated global
surface temperature anomalies are well within the uncertainty of reconstructions (medium confidence), except for some short periods
immediately following large volcanic eruptions, for which different forcing datasets disagree (Box TS.2, Figure 2b). {2.3.1.1, 3.3.3.1,
3.8.2.1, 7.4.4.1.2}
Proxy-based and model-simulated estimates of global surface temperature agree across multiple reference periods
(a) (b)
20
Global surface temperature relative to 1.0 Observed Reconstructed Simulated
1850–1900 0.8
15
Early Eocene
Simulated temperature (°C)
0.6
Global surface temperature
relative to 1850–1900 (°C)
10
0.4
5 0.2
mid-Pliocene
-
Recent past 0
0 1850–1900 Last Interglacial
TS mid-Holocene –0.2
–0.4
–5 Last Glacial Maximum
–0.6
–10
–10 –5 0 5 10 15 20 1000 1200 1400 1600 1800 2000
Reconstructed temperature (°C) Year (CE)
Box TS.2, Figure 2 | Global surface temperature as estimated from proxy records (reconstructed) and climate models (simulated). The intent of this
figure is to show the agreement between observations and models of global temperatures during paleo reference periods. (a) For individual paleoclimate reference
periods. (b) For the last millennium, with instrumental temperature (AR6 assessed mean, 10-year smoothed). Model uncertainties in (a) and (b) are 5–95% ranges of
multi-model ensemble means; reconstructed uncertainties are 5–95% ranges (medium confidence) of (a) midpoints and (b) multi-method ensemble median. {2.3.1.1,
Figure 2.34, Figure 3.2c, Figure 3.44}
Equilibrium climate sensitivity. Paleoclimate data provide evidence to estimate equilibrium climate sensitivity (ECS15) (Section
TS.3.2.1). In AR6, refinements in paleo data for paleoclimate reference periods indicate that ECS is very likely greater than 1.5°C and
likely less than 4.5°C, which is largely consistent with other lines of evidence and helps narrow the uncertainty range of the overall
assessment of ECS. Some of the CMIP6 climate models that have either high (>5°C) or low (<2°C) ECS also simulate past global
surface temperature changes outside the range of proxy-based reconstructions for the coldest and warmest reference periods. Since
AR5, independent lines of evidence, including proxy records from past warm periods and glacial–interglacial cycles, indicate that
sensitivity to forcing increases as temperature increases (Section TS.3.2.2). {7.4.3.2, 7.5.3, 7.5.6, Table 7.11}
Water cycle. New hydroclimate reconstructions and model-data comparisons have improved the understanding of the causes
and effects of long-term changes in atmospheric and ocean circulation, including monsoon variability and modes of variability
(Box TS.13, Section TS.4.2). Climate models are able to reproduce decadal drought variability on large regional scales, including the
severity, persistence and spatial extent of past megadroughts known from proxy records (medium confidence). Some long-standing
discrepancies remain, however, such as the magnitude of African monsoon precipitation during the early Holocene (the past 11,700
years), suggesting continuing knowledge gaps. Paleoclimate evidence shows that, in relatively high CO2 climates such as the Pliocene,
Walker circulation over the equatorial Pacific Ocean weakens, supporting the high confidence model projections of weakened Walker
cells by the end of the 21st century. {3.3.2, 8.3.1.6, 8.4.1.6, 8.5.2.1, 9.2}
15 In this Report, equilibrium climate sensitivity is defined as the equilibrium (steady state) change in the surface temperature following a doubling of the atmospheric carbon dioxide (CO2)
concentration from pre-industrial conditions.
46
Technical Summary
Sea level and ice sheets. Although past and future global warming differ in their forcings, evidence from paleoclimate
records and modelling show that ice-sheet mass and global mean sea level (GMSL) responded dynamically over multiple millennia
(high confidence). This evidence helps to constrain estimates of the committed GMSL response to global warming (Box TS.4). For
example, under a past global warming levels of around [2.5°C to 4°C] relative to 1850–1900, like during the mid-Pliocene Warm
Period, sea level was [5 to 25 m] higher than 1900 (medium confidence); under past global warming levels of [10°C to 18°C], like
during the Early Eocene, the planet was essentially ice free (high confidence). Constraints from these past warm periods, combined
with physical understanding, glaciology and modelling, indicate a committed long-term GMSL rise over 10,000 years, reaching about
8 to 13 m for sustained peak global warming of 2°C and up to 28 to 37 m for 5°C, which exceeds the AR5 estimate. {2.3.3.3, 9.4.1.4,
9.4.2.6, 9.6.2, 9.6.3.5}
Ocean. Since AR5, better integration of paleo-oceanographic data with modelling along with higher-resolution analyses of
transient changes have improved understanding of long-term ocean processes. Low-latitude sea surface temperatures at the Last
Glacial Maximum cooled more than previously inferred, resolving some inconsistencies noted in AR5. This paleo context supports
the assessment that ongoing increase in ocean heat content (OHC) represents a long-term commitment (see Core Concepts Box),
essentially irreversible on human time scales (high confidence). Estimates of past global OHC variations generally track those of
sea surface temperatures around Antarctica, underscoring the importance of Southern Ocean processes in regulating deep-ocean
temperatures. Paleoclimate data, along with other evidence of glacial–interglacial changes, show that Antarctic Circumpolar flow
strengthened and that ventilation of Antarctic Bottom Water accelerated during warming intervals, facilitating release of CO2 stored in
the deep ocean to the atmosphere. Paleo evidence suggests significant reduction of deep-ocean ventilation associated with meltwater
input during times of peak warmth. {2.3.1.1, 2.3.3.1, 9.2.2, 9.2.3.2} TS
Carbon cycle. Past climate states were associated with substantial differences in the inventories of the various carbon reservoirs,
including the atmosphere (Section TS.2.2). Since AR5, the quantification of carbon stocks has improved due to the development of
novel sedimentary proxies and stable-isotope analyses of air trapped in polar ice. Terrestrial carbon storage decreased markedly
during the Last Glacial Maximum by 300–600 PgC, possibly by 850 PgC when accounting for interactions with the lithosphere and
ocean sediments, a larger reduction than previously estimated, owing to a colder and drier climate. At the same time, the storage of
remineralized carbon in the ocean interior increased by as much as 750–950 PgC, sufficient to balance the removal of carbon from the
atmosphere (200 PgC) and terrestrial biosphere reservoirs combined (high confidence). {5.1.2.2}
TS.1.2 Progress in Climate Science satellite measurement techniques are now long enough to be
relevant for climate assessments. For example, globally distributed,
TS.1.2.1 Observation-based Products and their Assessments high-vertical-resolution profiles of temperature and humidity in the
upper troposphere and stratosphere can be obtained from the early
Observational capabilities have continued to improve and 2000s using global navigation satellite systems, leading to updated
expand overall since AR5, enabling improved consistency estimates of recent atmospheric warming. Improved measurements
between independent estimates of climate drivers, the of ocean heat content, warming of the land surface, ice-sheet mass
combined climate feedbacks, and the observed energy loss and sea level changes allow a better closure of the global
and sea level increase. Satellite climate records and energy and sea level budgets relative to AR5. For surface and
improved reanalyses are used as an additional line of balloon-based networks, apparent regional data reductions result
evidence for assessing changes at the global and regional from a combination of data policy issues, data curation/provision
scales. However, there have also been reductions in some challenges, and real cessation of observations, and are to an extent
observational data coverage or continuity and limited counter-balanced by improvements elsewhere. Limited observational
access to data resulting from data policy issues. Natural records of extreme events and spatial data gaps currently limit the
archives of past climate, such as tropical glaciers, have also assessment of some observed regional climate change. {1.5.1, 2.3.2,
been subject to losses (in part due to anthropogenic climate 7.2.2, Box 7.2, Cross-Chapter Box 9.1, 9.6.1, 10.2.2, 10.6, 11.2, 12.4}
change). {1.5.1, 1.5.2, 10.2.2}
New paleoclimate reconstructions from natural archives have enabled
Earth system observations are an essential driver of progress in our more robust reconstructions of the spatial and temporal patterns of
understanding of climate change. Overall, capabilities to observe past climate changes over multiple time scales (Box TS.2). However,
the physical climate system have continued to improve and expand. paleoclimate archives, such as tropical glaciers and modern natural
Improvements are particularly evident in ocean observing networks archives used for calibration (e.g., corals and trees), are rapidly
and remote-sensing systems. Records from several recently instigated disappearing owing to a host of pressures, including increasing
47
Technical Summary
temperatures (high confidence). Substantial quantities of past Section Box TS.1). Increasing horizontal resolution in global
instrumental observations of weather and other climate variables, climate models improves the representation of small-scale
over both land and ocean, which could fill gaps in existing datasets, features and the statistics of daily precipitation (high
remain un-digitized or inaccessible. These include measurements of confidence). Earth system models, which include additional
temperature (air and sea surface), rainfall, surface pressure, wind biogeochemical feedbacks, often perform as well as their
strength and direction, sunshine amount and many other variables lower-complexity global climate model counterparts, which
dating back into the 19th century. {1.5.1} do not account for these additional feedbacks (medium
confidence). {1.3.6, 1.5.3, 3.1, 3.5.1, 3.8.2, 4.3.1, 4.3.4, 7.5,
Reanalyses combine observations and models (e.g., a numerical 8.5.1, 9.6.3.1}
weather prediction model) using data assimilation techniques to
provide a spatially complete, dynamically consistent estimate of Climate model simulations coordinated and collected as part of the
multiple variables describing the evolving climate state. Since AR5, World Climate Research Programme’s Coupled Model Intercomparison
new reanalyses have been developed for the atmosphere and the Project Phase 6 (CMIP6), complemented by a range of results from
ocean with various combinations of increased resolution, extended the previous phase (CMIP5), constitute a key line of evidence
records, more consistent data assimilation and larger availability supporting this Report. The latest generation of CMIP6 models have
of uncertainty estimates. Limitations remain, for example, in how an improved representation of physical processes relative to previous
reanalyses represent global-scale changes to the water cycle. Regional generations, and a wider range of Earth system models now represent
reanalyses use high-resolution, limited-area models constrained by biogeochemical cycles. Higher-resolution models that better capture
regional observations and with boundary conditions from global smaller-scale processes are also increasingly becoming available
reanalyses. There is high confidence that regional reanalyses better for climate change research (Figure TS.2, Panels a and b). Results
represent the frequencies of extremes and variability in precipitation, from coordinated regional climate modelling initiatives, such as the
surface air temperature and surface wind than global reanalyses Coordinated Regional Climate Downscaling Experiment (CORDEX)
TS and provide estimates that are more consistent with independent complement and add value to the CMIP global models, particularly in
observations than dynamical downscaling approaches. {1.5.2, complex topography zones, coastal areas and small islands, as well
10.2.1.2, Annex I} as for extremes. {1.5.3, 1.5.4, 2.8.2, FAQ 3.3, 6.2.2, 6.4, 6.4.5, 8.5.1,
10.3.3, Atlas.1.4}
TS.1.2.2 Climate Model Performance
Projections of the increase in global surface temperature and the
This report assesses results from climate models pattern of warming from previous IPCC Assessment Reports and other
participating in the Coupled Model Intercomparison Project studies are broadly consistent with subsequent observations (limited
Phase 6 (CMIP6) of the World Climate Research Programme. evidence, high agreement), especially when accounting for the
These models include new and better representation of difference in radiative forcing scenarios used for making projections
physical, chemical and biological processes, as well as and the radiative forcings that actually occurred (Figure TS.3). The
higher resolution, compared to climate models considered AR5 and SROCC projections of GMSL for the 2007–2018 period have
in previous IPCC Assessment Reports. This has improved been shown to be consistent with observed trends in GMSL and
the simulation of the recent mean state of most large-scale regional weighted mean tide gauges. {1.3.6, 9.6.3.1}
indicators of climate change and many other aspects across
the climate system. Some differences from observations For most large-scale indicators of climate change, the simulated
remain, for example in regional precipitation patterns. recent mean climate from CMIP6 models underpinning this
Projections of the increase in global surface temperature, assessment have improved compared to the CMIP5 models used in
the pattern of warming, and global mean sea level rise AR5 (high confidence). This is evident from the performance of 18
from previous IPCC Assessment Reports and other studies simulated atmospheric and land large-scale indicators of climate
are broadly consistent with subsequent observations, change between the three generations of models (CMIP3, CMIP5,
especially when accounting for the difference in radiative and CMIP6) when benchmarked against reanalysis and observational
forcing scenarios used for making projections and the data (Figure TS.2, Panel c). Earth system models, characterized by
radiative forcings that actually occurred. additional biogeochemical feedbacks, often perform at least as well
as related, more constrained, lower-complexity models lacking these
The CMIP6 historical simulations assessed in this report feedbacks (medium confidence). {3.8.2, 10.3.3.3}
have an ensemble mean global surface temperature
change within 0.2°C of the observations over most of the The CMIP6 multi-model mean global surface temperature change
historical period, and observed warming is within the very from 1850–1900 to 2010–2019 is close to the best estimate of
likely range of the CMIP6 ensemble. However, some CMIP6 the observed warming. However, some CMIP6 models simulate
models simulate a warming that is either above or below a warming that is below or above the assessed very likely range. The
the assessed very likely range of observed warming. The CMIP6 models also reproduce surface temperature variations over
information about how well models simulate past warming, the past millennium, including the cooling that follows periods of
as well as other insights from observations and theory, are intense volcanism (medium confidence). For upper air temperature,
used to assess projections of global warming (see Cross- an overestimation of the upper tropical troposphere warming by
48
Technical Summary
about 0.1°C per decade between 1979 and 2014 persists in most evolution of the satellite-observed Arctic sea ice loss (high confidence).
CMIP5 and CMIP6 models (medium confidence), whereas the The ability to model ice-sheet processes has improved substantially
differences between simulated and improved satellite-derived since AR5. As a consequence, there is medium confidence in the
estimates of change in global mean temperature through the depth representation of key processes related to surface-mass balance and
of the stratosphere have decreased. {3.3.1} retreat of the grounding-line (the junction between a grounded ice
sheet and an ice shelf, where the ice starts to float) in the absence
Some CMIP6 models demonstrate an improvement in how clouds of instabilities. However, there remains low confidence in simulations
are represented. CMIP5 models commonly displayed a negative of ice-sheet instabilities, ice-shelf disintegration and basal melting
shortwave cloud radiative effect that was too weak in the present owing to their high sensitivity to both uncertain oceanic forcing and
climate. These errors have been reduced, especially over the uncertain boundary conditions and parameters. {1.5.3, 2.3.2, 3.4.1,
Southern Ocean, due to a more realistic simulation of supercooled 3.4.2, 3.8.2, 9.3.1, 9.3.2, 9.4.1, 9.4.2}
liquid droplets with sufficient numbers and an associated increase
in the cloud optical depth. Because a negative cloud optical depth CMIP6 models are able to reproduce most aspects of the spatial
feedback in response to surface warming results from ‘brightening’ structure and variance of the El Niño–Southern Oscillation (ENSO)
of clouds via active phase change from ice to liquid cloud particles and Indian Ocean Basin and Dipole modes of variability (medium
(increasing their shortwave cloud radiative effect), the extratropical confidence). However, despite a slight improvement in CMIP6, some
cloud shortwave feedback in CMIP6 models tends to be less negative, underlying processes are still poorly represented. Models reproduce
leading to a better agreement with observational estimates (medium observed spatial features and variance of the Southern Annular
confidence). CMIP6 models generally represent more processes Mode (SAM) and Northern Annular Mode (NAM) very well (high
that drive aerosol–cloud interactions than the previous generation confidence). The summertime SAM trend is well captured, with
of climate models, but there is only medium confidence that those CMIP6 models outperforming CMIP5 models (medium confidence).
enhancements improve their fitness-for-purpose of simulating By contrast, the cause of the NAM trend towards its positive phase
radiative forcing of aerosol–cloud interactions. {6.4, 7.4.2, FAQ 7.2} is not well understood. In the Tropical Atlantic basin, which contains TS
the Atlantic Zonal and Meridional modes, major biases in modelled
CMIP6 models still have deficiencies in simulating precipitation mean state and variability remain. Model performance is limited in
patterns, particularly in the tropical ocean. Increasing horizontal reproducing sea surface temperature anomalies for decadal modes
resolution in global climate models improves the representation of of variability, despite improvements from CMIP5 to CMIP6 (medium
small-scale features and the statistics of daily precipitation (high confidence) (see also Section TS.1.4.2.2, Table TS.4). {3.7.3–3.7.7}
confidence). There is high confidence that high-resolution global,
regional and hydrological models provide a better representation of Earth system models (ESMs) simulate globally averaged land
land surfaces, including topography, vegetation and land-use change, carbon sinks within the range of observation-based estimates
which can improve the accuracy of simulations of regional changes in (high confidence), but global-scale agreement masks large regional
the terrestrial water cycle. {3.3.2, 8.5.1, 10.3.3, 11.2.3} disagreements. There is also high confidence that the ESMs simulate
the weakening of the global net flux of CO2 into the ocean during
There is high confidence that climate models can reproduce the recent the 1990s, as well as the strengthening of the flux from 2000. {3.6}
observed mean state and overall warming of temperature extremes
globally and in most regions, although the magnitude of the trends Two important quantities used to estimate how the climate system
may differ. There is high confidence in the ability of models to capture responds to changes in greenhouse gas (GHG) concentrations are the
the large-scale spatial distribution of precipitation extremes over land. equilibrium climate sensitivity (ECS) and transient climate response
The overall performance of CMIP6 models in simulating the intensity (TCR16). The CMIP6 ensemble has broader ranges of ECS and TCR
and frequency of extreme precipitation is similar to that of CMIP5 values than CMIP5 (see Section TS.3.2 for the assessed range). These
models (high confidence). {Cross-Chapter Box 3.2, 11.3.3, 11.4.3} higher sensitivity values can, in some models, be traced to changes
in extratropical cloud feedbacks (medium confidence). To combine
The structure and magnitude of multi-model mean ocean temperature evidence from CMIP6 models and independent assessments of ECS
biases have not changed substantially between CMIP5 and CMIP6 and TCR, various emulators are used throughout the report. Emulators
(medium confidence). Since AR5, there is improved consistency are a broad class of simple climate models or statistical methods
between recent observed estimates and model simulations of changes that reproduce the behaviour of complex ESMs to represent key
in upper (<700 m) ocean heat content. The mean zonal and overturning characteristics of the climate system, such as global surface temperature
circulations of the Southern Ocean and the mean overturning and sea level projections. The main application of emulators in AR6
circulation of the North Atlantic (AMOC) are broadly reproduced by is to extrapolate insights from ESMs and observational constraints to
CMIP5 and CMIP6 models. {3.5.1, 3.5.4, 9.2.3, 9.3.2, 9.4.2} produce projections from a larger set of emissions scenarios, which
is achieved due to their computational efficiency. These emulated
CMIP6 models better simulate the sensitivity of Arctic sea ice area projections are also used for scenario classification in WGIII. {Box 4.1,
to anthropogenic CO2 emissions, and thus better capture the time 4.3.4, 7.4.2, 7.5.6, Cross-Chapter Box 7.1, FAQ 7.2}
16 In this Report, transient climate response is defined as the surface temperature response for the hypothetical scenario in which atmospheric carbon dioxide (CO2) increases at 1% yr –1 from
pre-industrial to the time of a doubling of atmospheric CO2 concentration.
49
Technical Summary
Number of models
150 70
60 30
Level
200
km
50
250 40 20
300 30
20 10
350
10
0
400 0
Atm Res Oce Res Atm lev Ocn lev # Mod Aerosols Atm Chem Land carbon N cycle vegetation Ocean BGC
cycle prognostic
(c) Pattern correlation with observational reference
1.0
0.8
Correlation
0.4
CMIP6
0.2 CMIP5
CMIP3
Additional observations
TS
0
Near-Surface Precipitation TOA TOA TOA SW TOA LW Sea Level Temperature Temperature Eastward Eastward Northward Northward Geopotential Specific Soil Leaf Area Gross Primary
Air Temperature Outgoing Outgoing Cloud Rad Cloud Rad Pressure 850 hPa 200 hPa Wind Wind Wind Wind Height Humidity Moisture Index Productivity
Shortwave Longwave Effect Effect 850 hPa 200 hPa 850 hPa 200 hPa 500 hPa 400 hPa
Radiation Radiation
Figure TS.2 | Progress in climate models. The intent of this figure is to show present improvements in climate models in resolution, complexity and representation of key
variables. (a) Evolution of model horizontal resolution and vertical levels (based on Figure 1.19). (b) Evolution of inclusion of processes and resolution from Coupled Model
Intercomparison Project Phase 3 (CMIP3), Phase 5 (CMIP5) and Phase 6 (CMIP6; Annex II). (c) Centred pattern correlations between models and observations for the annual
mean climatology over the period 1980–1999. Results are shown for individual CMIP3 (cyan), CMIP5 (blue) and CMIP6 (red) models (one ensemble member is used) as short
lines, along with the corresponding ensemble averages (long lines). The correlations are shown between the models and the primary reference observational data set (from left
to right: ERA5, GPCP-SG, CERES-EBAF, CERES-EBAF, CERES-EBAF, CERES-EBAF, JR-55, ERA5, ERA5, ERA5, ERA5, ERA5, ERA5, AIRS, ERA5, ESACCI-Soilmoisture, LAI3g, MTE). In
addition, the correlation between the primary reference and additional observational data sets (from left to right: NCEP, GHCN, -, -, -, -, ERA5, HadISST, NCEP, NCEP, NCEP, NCEP,
NCEP, NCEP, ERA5, NCEP, -, -, FLUXCOM) are shown (solid grey circles) if available. To ensure a fair comparison across a range of model resolutions, the pattern correlations
are computed after regridding all datasets to a resolution of 4º in longitude and 5º in latitude. (Expanded from Figure 3.43; produced with ESMValTool version 2). {Figure 3.43}
TS.1.2.3 Understanding Climate Variability and Observational datasets have been extended and improved since
Emerging Changes AR5, providing stronger evidence that the climate is changing and
allowing better estimates of natural climate variability on decadal
Observed changes in climate are unequivocal at the global time scales. There is very high confidence that the slower rate of global
scale and are increasingly apparent on regional and local surface temperature change observed over 1998–2012 compared to
spatial scales. Both the rate of long-term change and 1951–2012 was temporary, and was, with high confidence, induced
the amplitude of year-to-year variations differ between by internal variability (particularly Pacific Decadal Variability) and
regions and across climate variables, thus influencing when variations in solar irradiance and volcanic forcing that partly offset the
changes emerge or become apparent compared to natural anthropogenic warming over this period. Global ocean heat content
variations (see Emergence in Core Concepts Box). The signal continued to increase throughout this period, indicating continuous
of temperature change has emerged more clearly in tropical warming of the entire climate system (very high confidence). Hot
regions, where year-to-year variations tend to be small over extremes also continued to increase during this period over land
land, than in regions with greater warming but larger year- (high confidence). Even in a continually warming climate, periods
to-year variations (high confidence) (Figure TS.3). Long- of reduced and increased trends in global surface temperature at
term changes in other variables have emerged in many decadal time scales will continue to occur in the 21st century (very
regions, such as for some weather and climate extremes high confidence). {Cross-Chapter Box 3.1, 3.3.1, 3.5.1, 4.6.2, 11.3.2}
and Arctic sea ice area. {1.4.2, Cross-Chapter Box 3.1, 9.3.1,
11.3.2, 12.5.2} Since AR5, the increased use of ‘large ensembles’, or multiple
simulations with the same climate model but using different initial
conditions, supports improved understanding of the relative roles
50
Technical Summary
of internal variability and forced change in the climate system. occurred, and there is medium confidence that many of these changes
Simulations and understanding of modes of climate variability, are attributable to human activities. Several impact-relevant changes
including teleconnections, have improved since AR5 (medium have not yet emerged from natural variability but will emerge sooner
confidence), and larger ensembles allow a better quantification of or later in this century depending on the emissions scenario (high
uncertainty in projections due to internal climate variability. {1.4.2, confidence). Ocean acidification and deoxygenation have already
1.5.3, 1.5.4, 4.2, 4.4.1, Box 4.1, 8.5.2, 10.3.4, 10.4} emerged over most of the global open ocean, as has a reduction in
Arctic sea ice (high confidence). {9.3.1, 9.6.4, 11.2, 11.3, 12.4, 12.5,
Changes in regional climate can be detected even though natural Atlas.3–Atlas.11}
climate variations can temporarily increase or obscure anthropogenic
climate change on decadal time scales. While anthropogenic forcing TS.1.2.4 Understanding of Human Influence
has contributed to multi-decadal mean precipitation changes in
several regions, internal variability can delay emergence of the The evidence for human influence on recent climate change
anthropogenic signal in long-term precipitation changes in many has strengthened progressively from the IPCC Second
land regions (high confidence). {10.4} Assessment Report to AR5 and is even stronger in this
assessment, including for regional scales and for extremes.
Mean temperatures and heat extremes have emerged above natural Human influence in the IPCC context refers to the human
variability in almost all land regions with high confidence. Changes activities that lead to or contribute to a climate response,
in temperature-related variables, such as regional temperatures, such as the human-induced emissions of greenhouse
growing season length, extreme heat and frost, have already gases that subsequently alter the atmosphere’s radiative
30°S
60°S
Missing
90°S
data 0.5 1 2 3 4 5
–2.4 –2.1 –1.8 –1.5 –2.1 –0.9 –0.6 –0.3 0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4
°C
(b) Change in temperature at a global warming level of 1°C relative to the size of year-to-year variations
Zonal mean
90°N
60°N
30°N
30°S
60°S
Missing 90°S
data 0.5 1 2 3 4 5
Figure TS.3 | Emergence of changes in temperature over the historical period. The intent of this figure is to show how observed changes in temperature have
emerged and that the emergence pattern agrees with model simulations. The observed change in temperature at a global warming level of 1°C (a), and the signal-to-noise ratio
(the change in temperature at a global warming level of 1°C, divided by the size of year-to-year variations, (b)) using data from Berkeley Earth. The right panels show the zonal
means of the maps and include data from different observational datasets (red) and the Coupled Model Intercomparison Project Phase 6 (CMIP6) simulations (black, including
the 5–95% range) processed in the same way as the observations. {1.4.2, 10.4.3}
51
Technical Summary
properties, resulting in warming of the atmosphere, ocean of scientific literature combining different lines of evidence, and
and land components of the climate system. Other human improved accessibility to different types of climate models (high
activities influencing climate include the emission of confidence) (see Sections TS.2 and TS.4). {Cross-Working Group Box:
aerosols and other short-lived climate forcers, and land-use Attribution in Chapter 1, 1.5, 3.2, 3.5, 5.2, 6.4.3, 8.3, 9.6, 10.1, 10.2,
change such as urbanization. Progress in our understanding 10.3.3, 10.4.1, 10.4.2, 10.4.3, 10.5, 10.6, Cross-Chapter Box 10.3,
of human influence is gained from longer observational Box 10.3, 11.1.6, 11.2–11.9, 12.4}
datasets, improved paleoclimate information, a stronger
warming signal since AR5, and improvements in climate
models, physical understanding and attribution techniques TS.1.3 Assessing Future Climate Change
(see Core Concepts Box). Since AR5, the attribution to
human influence has become possible across a wider range Various frameworks can be used to assess future climatic changes and
of climate variables and climatic impact-drivers (CIDs, to synthesize knowledge across climate change assessment in WGI,
see Core Concepts Box). New techniques and analyses WGII and WGIII. These frameworks include: (i) scenarios, (ii) global
drawing on several lines of evidence have provided greater warming levels and (iii) cumulative CO2 emissions (see Core Concepts
confidence in attributing changes in regional weather and Box). The latter two offer scenario- and path-independent approaches
climate extremes to human influence (high confidence). to assess future projections. Additional choices, for instance with regard
{1.3, 1.5.1, Appendix 1.A, 3.1–3.8, 5.2, 6.4.2, 7.3.5, 7.4.4, to common reference periods and time windows for which changes
8.3.1, 10.4, Cross-Chapter Box 10.3, 11.2–11.9, 12.4} are assessed, can further help to facilitate integration across the WGI
report and across the whole AR6 (see Section TS.1.1). {1.4.1, 1.6, Cross-
Combining the evidence from across the climate system increases Chapter Box 1.4, 4.2.2, 4.2.4, Cross-Chapter Box 11.1}
the level of confidence in the attribution of observed climate change
to human influence and reduces the uncertainties associated with TS.1.3.1 Climate Change Scenarios
TS assessments based on single variables. {Cross-Chapter Box 10.3}
A core set of five illustrative scenarios based on the Shared
Since AR5, the accumulation of energy in the Earth system has Socio-economic Pathways (SSPs) are used consistently
become established as a robust measure of the rate of global across this Report: SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0,
climate change on interannual-to-decadal time scales. The rate of and SSP5-8.5. These scenarios cover a broader range of
accumulation of energy is equivalent to Earth’s energy imbalance greenhouse gas and air pollutant futures than assessed in
and can be quantified by changes in the global energy inventory for earlier WGI reports, and they include high-CO2 emissions
all components of the climate system, including global ocean heat pathways without climate change mitigation as well as new
uptake, warming of the atmosphere, warming of the land and melting low-CO2 emissions pathways (Figure TS.4). In these scenarios,
of ice. Compared to changes in global surface temperature, Earth’s differences in air pollution control and variations in climate
energy imbalance (see Core Concepts Box) exhibits less variability, change mitigation stringency strongly affect anthropogenic
enabling more accurate identification and estimation of trends. {Box emissions trajectories of SLCFs. Modelling studies relying
7.2 and Section 7.2} on the Representative Concentration Pathways (RCPs) used
in AR5 complement the assessment based on SSP scenarios,
Identifying the human-induced components contributing to the for example at the regional scale.
energy budget provides an implicit estimate of the human influence
on global climate change (Sections TS.2 and TS.3.1). {Cross-Working A comparison of simulations from CMIP5 using the RCPs
Group Box: Attribution in Chapter 1, 3.8, 7.2.2, Box 7.2, with SSP-based simulations from CMIP6 shows that about
Cross-Chapter Box 9.1} half of the increase in simulated warming in CMIP6 versus
CMIP5 arises because higher climate sensitivity is more
Regional climate changes can be moderated or amplified by regional prevalent in CMIP6 model versions; the other half arises
forcing from land-use and land-cover changes or from aerosol from higher radiative forcing in nominally corresponding
concentrations and other short-lived climate forcers (SLCFs). For scenarios (e.g., RCP8.5 and SSP5-8.5; medium confidence).
example, the difference in observed warming trends between cities The feasibility or likelihood of individual scenarios is not
and their surroundings can partly be attributed to urbanization part of this assessment, which focuses on the climate
(very high confidence). While established attribution techniques response to a large range of emissions scenarios. {1.5.4, 1.6,
provide confidence in our assessment of human influence on Cross-Chapter Box 1.4, 4.2, 4.3, 4.6, 6.6, 6.7, Cross-Chapter
large-scale climate changes (as described in Section TS.2), new Box 7.1, Atlas.2.1}
techniques developed since AR5, including attribution of individual
events, have provided greater confidence in attributing changes in Climate change projections with climate models require information
climate extremes to climate change (Box TS.10). Multiple attribution about future emissions or concentrations of greenhouse gases,
approaches support the contribution of human influence to several aerosols, ozone-depleting substances, and land use over time
regional multi-decadal mean precipitation changes (high confidence). (Figure TS.4). This information can be provided by scenarios, which
Understanding about past and future changes in weather and climate are internally consistent projections of these quantities based on
extremes has increased due to better observation-based datasets, assumptions of how socio-economic systems could evolve over the
physical understanding of processes, an increasing proportion 21st century. Emissions from natural sources, such as the ocean and
52
Technical Summary
Human
activities
Other
CO2 (e.g., SLCF,
land use
albedo)
Non-CO2
Greenhouse
gases
Emissions N2O
HFCs etc.
200
NH3
Land use albedo etc.
900
175 NOx
125
CO2 800 CH4
100 700 150
MtNOx / yr
600 125
MtCH4 / yr
GtCO2 / yr
75 500 100
50 400
75
25 300
200 50
0 100 25
-25 0 0
1950 2000 2050 2100 1950 2000 2050 2100 1950 2000 2050 2100
Concentrations 4000
N2O
HFCs etc.
140
120
1000
CO2 3500 CH4 100
MtSO2 / yr
3000 80
800
2500 TS
ppm
60
ppb
600 2000 40
400 1500 20 SO2
1000 0
200 1950 2000 2050 2100
1950 2000 2050 2100 1950 2000 2050 2100
12
Radiative
Effective radiative forcing
10
Forcing 8
6
Total anthropogenic
(W/m-2)
4
Carbon- 2
Cycle and 0
non-CO2 Natural
-2
biogeo- 1950 2000 2050 2100
chemical
feedbacks
7
Global
(°C rel. to 1850–1900)
Change in global surface
6
Warming 5 Projections
temperature
4
3
Legend:
2
Historical Observations
1
SSP5-8.5
SSP3-7.0 0
1950 2000 2050 2100
SSP2-4.5 2100
SSP1-2.6 RCP
SSP1-1.9 range Regional Climate Change
Temperature Precipitation
GWL 2°C GWL 2°C
Climatic
Impact-
Robust significant change
No or no robust -5 -4 -3 -2.5 -2 -1.5 0 1.5 2 2.5 3 4 5
Drivers -40 -20 0 +20 +40
Legend
see left
significant change Change in annual mean
Change in annual mean
Conflicting signal surface temperature (°C) precipitation (%)
Figure TS.4 | The climate change cause–effect chain: The intent of this figure is to illustrate the process chain starting from anthropogenic emissions, to changes in
atmospheric concentration, to changes in Earth’s energy balance (‘forcing’), to changes in global climate and ultimately regional climate and climatic impact-drivers. Shown
is the core set of five Shared Socio-economic Pathway (SSP) scenarios as well as emissions and concentration ranges for the previous Representative Concentration Pathway
(RCP) scenarios in year 2100; carbon dioxide (CO2) emissions (GtCO2 yr–1), panel top left; methane (CH4) emissions (middle) and sulphur dioxide (SO2), nitrogen oxide (NOx)
emissions (all in Mt yr–1), top right; concentrations of atmospheric CO2 (ppm) and CH4 (ppb), second row left and right; effective radiative forcing for both anthropogenic and
natural forcings (W m–2), third row; changes in global surface air temperature (°C) relative to 1850–1900, fourth row; maps of projected temperature change (°C) (left) and
changes in annual-mean precipitation (%) (right) at a global warming level (GWL) of 2°C relative to 1850–1900 (see also Figure TS.5), bottom row. Carbon cycle and non-CO2
biogeochemical feedbacks will also influence the ultimate response to anthropogenic emissions (arrows on the left). {1.6.1, Cross-Chapter Box 1.4, 4.2.2, 4.3.1, 4.6.1, 4.6.2}
53
Technical Summary
the land biosphere, are usually assumed to be constant, or to evolve at the regional scale (Section TS.4). Scenario extensions are based
in response to changes in anthropogenic forcings or to projected on assumptions about the post-2100 evolution of emissions or of
climate change. Natural forcings, such as past changes in solar radiative forcing that are independent from the modelling of socio-
irradiance and historical volcanic eruptions, are represented in model economic dynamics, which does not extend beyond 2100. To explore
simulations covering the historical era. Future simulations assessed specific dimensions, such as air pollution or temporary overshoot of
in this Report account for projected changes in solar irradiance and a given warming level, scenario variants are used in addition to the
for the long-term mean background forcing from volcanoes, but not core set. {1.6.1, Cross-Chapter Box 1.4, 4.2.2, 4.2.6, 4.7.1, Cross-
for individual volcanic eruptions. Scenarios have a long history in Chapter Box 7.1}
IPCC as a method for systematically examining possible futures and
following the cause–effect chain: from anthropogenic emissions, to SSP1-1.9 represents the low end of future emissions pathways,
changes in atmospheric concentrations, to changes in Earth’s energy leading to warming below 1.5°C in 2100 and limited temperature
balance (‘forcing’), to changes in global climate and ultimately overshoot of 1.5°C over the course of the 21st century (see
regional climate and climatic impact-drivers (Figure TS.4, Section Figure TS.6). At the opposite end of the range, SSP5-8.5 represents
TS.2, Infographic TS.1). {1.5.4, 1.6.1, 4.2.2, 4.4.4, Cross-Chapter the very high warming end of future emissions pathways from the
Box 4.1, 11.1} literature. SSP3-7.0 has overall lower GHG emissions than SSP5-8.5
but, for example, CO2 emissions still almost double by 2100 compared
The uncertainty in climate change projections that results from to today’s levels. SSP2-4.5 and SSP1-2.6 represent scenarios with
assessing alternative socio-economic futures, the so-called scenario stronger climate change mitigation and thus lower GHG emissions.
uncertainty, is explored through the use of scenario sets. Designed SSP1-2.6 was designed to limit warming to below 2°C. Infographic
to span a wide range of possible future conditions, these scenarios TS.1 presents a narrative depiction of SSP-related climate futures.
do not intend to match how events actually unfold in the future, No likelihood is attached to the scenarios assessed in this Report,
and they do not account for impacts of climate change on the socio- and the feasibility of specific scenarios in relation to current trends
TS economic pathways. Besides scenario uncertainty, climate change is best informed by the WGIII contribution to AR6. In the scenario
projections are also subject to climate response uncertainty (i.e., the literature, the plausibility of some scenarios with high CO2 emissions,
uncertainty related to our understanding of the key physical processes such as RCP8.5 or SSP5-8.5, has been debated in light of recent
and structural uncertainties in climate models) and irreducible and developments in the energy sector. However, climate projections from
intrinsic uncertainties related to internal variability. Depending on these scenarios can still be valuable because the concentration levels
the spatial and temporal scales of the projection, and on the variable reached in RCP8.5 or SSP5-8.5 and corresponding simulated climate
of interest, the relative importance of these different uncertainties futures cannot be ruled out. That is because of uncertainty in carbon-
may vary substantially. {1.4.3, 1.6, 4.2.5, Box 4.1, 8.5.1} cycle feedbacks which, in nominally lower emissions trajectories, can
result in projected concentrations that are higher than the central
Scenarios in AR6 cover a broader range of emissions futures than concentration levels typically used to drive model projections. {1.6.1;
considered in AR5, including high CO2 emissions scenarios without Cross-Chapter Box 1.4; 4.2.2, 5.4; SROCC; Chapter 3 in WGIII}
climate change mitigation as well as a low CO2 emissions scenario
reaching net zero CO2 emissions (see Core Concepts Box) around mid- The socio-economic narratives underlying SSP-based scenarios
century. In this Report, a core set of five illustrative scenarios is used differ in their assumed level of air pollution control. Together with
to explore climate change over the 21st century and beyond (Section variations in climate change mitigation stringency, this difference
TS.2). They are labelled SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0, and strongly affects anthropogenic emissions trajectories of SLCFs,
SSP5-8.517 and span a wide range of radiative forcing levels in 2100. some of which are also air pollutants. SSP1 and SSP5 assume strong
They start in 2015 and include scenarios with high and very high pollution control, projecting a decline of global emissions of ozone
GHG emissions and CO2 emissions that roughly double from current precursors (except methane; CH4) and of aerosols and most of
levels by 2100 and 2050, respectively (SSP3-7.0 and SSP5-8.5); their precursors in the mid- to long term. The reductions due to air
scenarios with intermediate GHG emissions and CO2 emissions pollution controls are further strengthened in scenarios that assume
remaining around current levels until the middle of the century a marked decarbonization, such as SSP1-1.9 or SSP1-2.6. SSP2-4.5
(SSP2-4.5); and scenarios with very low and low GHG emissions and is a medium pollution-control scenario with air pollutant emissions
CO2 emissions declining to net zero around or after 2050, followed by following current trends, and SSP3-7.0 is a weak pollution-control
varying levels of net negative CO2 emissions (SSP1-1.9 and SSP1-2.6). scenario with strong increases in emissions of air pollutants over the
These SSP scenarios offer unprecedented detail of input data for ESM 21st century. Methane emissions in SSP-based scenarios vary with
simulations and allow for a more comprehensive assessment of the overall climate change mitigation stringency, declining rapidly
climate drivers and responses, in particular because some aspects, in SSP1-1.9 and SSP1-2.6 but declining only after 2070 in SSP5-8.5.
such as the temporal evolution of pollutants, emissions or changes in SSP trajectories span a wider range of air pollutant emissions than
land use and land cover, span a broader range in the SSP scenarios considered in the RCP scenarios (see Figure TS.4), reflecting the
than in the RCPs used in AR5. Modelling studies utilizing the RCPs potential for large regional differences in their assumed pollution
complement the assessment based on SSP scenarios, for example,
17 Throughout this Report, scenarios are referred to as SSPx-y, where “SSPx” refers to the Shared Socio-economic Pathway or “SSP” describing the socio-economic trends underlying the scenario,
and “y” refers to the approximate target level of radiative forcing (in W m–-2) resulting from the scenario in the year 2100.
54
Technical Summary
policies. Their effects on climate and air pollution are assessed in Box TS.1.3.2 Global Warming Levels and Cumulative CO2 Emissions
TS.7. {4.4.4, 6.6.1, Figure 6.4, 6.7.1, Figure 6.19}
Quantifying geographical response patterns of climate
Since the RCPs are also labelled by the level of radiative forcing change at various global warming levels (GWLs), such as
they reach in 2100, they can in principle be related to the core set 1.5°C or 2°C above the 1850–1900 period, is useful for
of AR6 scenarios (Figure TS.4). However, the RCPs and SSP-based characterizing changes in mean climate, extremes and
scenarios are not directly comparable. First, the gas-to-gas climatic impact-drivers. Global warming levels are used
compositions differ; for example, the SSP5-8.5 scenario has higher in this Report as a dimension of integration independent
CO2 but lower CH4 concentrations compared to RCP8.5. Second, the of the timing when the warming level is reached and of
projected 21st-century trajectories may differ, even if they result the emissions scenario that led to the warming. For many
in the same radiative forcing by 2100. Third, the overall effective climate variables the response pattern for a given GWL is
radiative forcing (see Core Concepts Box) may differ, and tends consistent across different scenarios. However, this is not
to be higher for the SSPs compared to RCPs that share the same the case for slowly responding processes, such as ice-sheet
nominal stratospheric-temperature-adjusted radiative forcing label. and glacier mass loss, deep ocean warming, and the related
Comparing the differences between CMIP5 and CMIP6 projections sea level rise. The response of these variables depends on
(Cross-Section Box TS.1) that were driven by RCPs and SSP-based the time it takes to reach the GWL, differs if the warming is
scenarios, respectively, indicates that about half of the difference in reached in a transient warming state or after a temporary
simulated warming arises because of higher climate sensitivity being overshoot of the warming level, and will continue to evolve,
more prevalent in CMIP6 model versions; the remainder arises from over centuries to millennia, even after global warming
higher ERF in nominally corresponding scenarios (e.g., RCP8.5 and has stabilized. Different GWLs correspond closely to
SSP5-8.5; medium confidence) (see Section TS.1.2.2). In SSP1-2.6 and specific cumulative CO2 emissions due to their near-linear
SSP2-4.5, changes in ERF also explain about half of the changes in the relationship with global surface temperature. This Report
range of warming (medium confidence). For SSP5-8.5, higher climate uses 1.0°C, 1.5°C, 2.0°C, 3.0°C and 4.0°C above 1850–1900 TS
sensitivity is the primary reason behind the upper end of the CMIP6- conditions as a primary set of GWLs. {1.6.2, 4.2.4, 4.6.1, 5.5,
projected warming being higher than for RCP8.5 in CMIP5 (medium Cross-Chapter Box 11.1, Cross-chapter Box 12.1}
confidence). Note that AR6 uses multiple lines of evidence beyond
CMIP6 results to assess global surface temperature under various For many indicators of climate change, such as seasonal and annual
scenarios (see Cross-Section Box TS.1 for the detailed assessment). mean and extreme surface air temperatures and precipitation, the
{1.6, 4.2.2, 4.6.2.2, Cross-Chapter Box 7.1} geographical patterns of changes are well estimated by the level of
global surface warming, independently of the details of the emissions
Earth system models can be driven by anthropogenic CO2 pathways that caused the warming, or the time at which the level of
emissions (‘emissions-driven’ runs), in which case atmospheric warming is attained. GWLs, defined as a global surface temperature
CO2 concentration is a projected variable; or by prescribed time- increase of, for example, 1.5°C or 2°C relative to the mean of 1850–
varying atmospheric concentrations (‘concentration-driven’ runs). In 1900, are therefore a useful way to integrate climate information
emissions-driven runs, changes in climate feed back on the carbon independently of specific scenarios or time periods. {1.6.2, 4.2.4,
cycle and interactively modify the projected CO2 concentration 4.6.1, 11.2.4, Cross-Chapter Box 11.1}
in each ESM, thus adding the uncertainty in the carbon cycle
response to climate change to the projections. Concentration- The use of GWLs allows disentangling the contribution of changes
driven simulations are based on a central estimate of carbon cycle in global warming from regional aspects of the climate response,
feedbacks, while emissions-driven simulations help quantify the role as scenario differences in response patterns at a given GWL are
of feedback uncertainty. The differences in the few ESMs for which often smaller than model uncertainty and internal variability. The
both emissions and concentration-driven runs were available for the relationship between the GWL and response patterns is often linear,
same scenario are small and do not affect the assessment of global but integration of information can also be done for non-linear
surface temperature projections discussed in Cross-Section Box TS.1 changes, like the frequency of heat extremes. The requirement is that
and Section TS.2 (high confidence). By the end of the 21st century, the relationship to the GWL is broadly independent of the scenario
emissions-driven simulations are on average around 0.1°C cooler and relative contribution of radiative forcing agents. {1.6, 11.2.4,
than concentration-driven runs, reflecting the generally lower CO2 Cross-Chapter Box 11.1}
concentrations simulated by the emissions-driven ESMs, and have a
spread about 0.1°C greater, reflecting the range of simulated CO2 The GWL approach to integration of climate information also has
concentrations. However, these carbon cycle–climate feedbacks do some limitations. Variables that are quick to respond to warming,
affect the transient climate response to cumulative CO2 emissions like temperature and precipitation, including extremes, sea ice area,
(TCRE18), and their quantification is crucial for the assessment of permafrost and snow cover, show little scenario dependence for a
remaining carbon budgets consistent with global warming levels given GWL, whereas slow-responding variables such as glacier and
simulated by ESMs (see Section TS.3). {1.6.1, Cross-Chapter Box 1.4, ice-sheet mass, warming of the deep ocean and their contributions
4.2, 4.3.1, 5.4.5, Cross-Chapter Box 7.1} to sea level rise, have substantial dependency on the trajectory of
18 The transient surface temperature change per unit of cumulative CO2 emissions, usually 1000 GtC.
55
Technical Summary
warming taken to reach the GWL. A given GWL can also be reached SR1.5 concluded that ‘climate models project robust differences in
for different balances between anthropogenic forcing agents, such regional climate characteristics between present-day and global
as long-lived greenhouse gas and SLCF emissions, and the response warming of 1.5°C, and between 1.5°C and 2°C’. This Report adopts
patterns may depend on this balance. Finally, there is a difference a set of common GWLs across which climate projections, impacts,
in the response even for temperature-related variables if a GWL is adaptation challenges and climate change mitigation challenges can
reached in a rapidly warming transient state or in an equilibrium be integrated, within and across the three Working Groups, relative
state when the land–sea warming contrast is less pronounced. In this to 1850–1900. The core set of GWLs in this Report are 1.0°C (close
Report, the climate responses at different GWLs are calculated based to present day conditions), 1.5°C, 2.0°C, 3.0°C and 4.0°C. {1.4, 1.6.2,
on climate model projections for the 21st century (see Figure TS.5), Cross-Chapter Box 1.2, Table 1.5, Cross-Chapter Box 11.1}
which are mostly not in equilibrium. The SSP1-1.9 scenario allows
assessing the response to a GWL of about 1.5°C after a (relatively) Connecting Scenarios and Global Warming Levels
short-term stabilization by the end of the 21st century. {4.6.2, 9.3.1.1,
9.5.2.3, 9.5.3.3, 11.2.4, Cross-Chapter Box 11.1, Cross-Chapter In this Report, scenario-based climate projections are translated
Box 12.1} into GWLs by aggregating the ESM model response at specific
GWLs across scenarios (see Figure TS.5 and Figure TS.6). The climate
Global warming levels are highly relevant as a dimension of response pattern for the 20-year period around when individual
integration across scientific disciplines and socio-economic actors simulations reach a given GWL are averaged across all models and
and are motivated by the long-term goal in the Paris Agreement scenarios that reach that GWL. The best estimate and likely range
of ‘holding the increase in the global average temperature to well of the timing of when a certain GWL is reached under a particular
below 2°C above pre-industrial levels and to pursue efforts to limit scenario (or ‘GWL-crossing time’), however, is based not only on
the temperature increase to 1.5°C above pre-industrial levels’. CMIP6 output, but on a combined assessment taking into account
The evolution of aggregated impacts with temperature levels has the observed warming to date, CMIP6 output and additional lines
TS also been widely used and embedded in the WGII assessment. of evidence (see Cross-Section Box TS.1). {4.3.4, Cross-Chapter
This includes the ‘Reasons for Concern’ (RFC) and other ‘burning Box 11.1, Atlas.2, Interactive Atlas}
ember’ diagrams in IPCC WGII. The RFC framework has been further
expanded in SR1.5, SROCC and SRCCL by explicitly looking at the Global warming levels are closely related to cumulative CO2 (and
differential impacts between half-degree GWLs and the evolution of in some cases CO2-equivalent) emissions. This Report confirms the
risk for different socio-economic assumptions. {1.4.4, 1.6.2, 11.2.4, assessment of the WGI contribution to AR5 and SR1.5 that a near-
12.5.2, Cross-Chapter Box 11.1, Cross-Chapter Box 12.1} linear relationship exists between cumulative CO2 emissions and the
(a) Global mean temperature in CMIP6 (b) Patterns of change in near-surface air temperature, precipitation and soil moisture
5 SSP3-7.0 (20-yr GSAT means) Temperature change Precipitation change Soil moisture change
SSP1-2.6 (20-yr GSAT means) 48 47 43
+4°C +4°C
4
+2°C
+2°C
2
+1.5°C 153 147 132
1 +1.5°C
2000 2020 2040 2060 2080 2100 -5 -4 -3-2.5-2 -1.5 0 1.5 2 2.5 3 4 5
°C
-40 -30 -20 -10 0 10 20 30 40
%
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
σ
Figure TS.5 | Scenarios, global warming levels, and patterns of change. The intent of this figure is to show how scenarios are linked to global warming levels (GWLs)
and to provide examples of the evolution of patterns of change with global warming levels. (a) Illustrative example of GWLs defined as global surface temperature response
to anthropogenic emissions in unconstrained Coupled Model Intercomparison Project Phase 6 (CMIP6) simulations, for two illustrative scenarios (SSP1-2.6 and SSP3-7.0). The
time when a given simulation reaches a GWL, for example, +2°C, relative to 1850–1900 is taken as the time when the central year of a 20-year running mean first reaches
that level of warming. See the dots for +2°C, and how not all simulations reach all levels of warming. The assessment of the timing when a GWL is reached takes into account
additional lines of evidence and is discussed in Cross-Section Box TS.1. (b) Multi-model, multi-simulation average response patterns of change in near-surface air temperature,
precipitation (expressed as percentage change) and soil moisture (expressed in standard deviations of interannual variability) for three GWLs. The number to the top right of
the panels shows the number of model simulations averaged across including all models that reach the corresponding GWL in any of the five Shared Socio-economic Pathways
(SSPs). See Section TS.2 for discussion. {Cross-Chapter Box 11.1}
56
Technical Summary
resulting increase in global surface temperature (Section TS.3.2). This as sub-continents and oceanic regions, or to typological regions, such
implies that continued CO2 emissions will cause further warming as monsoon regions, coastlines, mountain ranges or cities, as used in
and associated changes in all components of the climate system. For Section TS.4. A new set of standard AR6 WGI reference regions has
declining cumulative CO2 emissions (i.e., if negative net emissions also been included in this Report (Figure TS.6, bottom panels). {1.4.5,
are achieved), the relationship is less strong for some components, 10.1, 11.9, 12.1–12.4, Atlas.1.3.3–1.3.4}
such as the hydrological cycle. The WGI report uses cumulative CO2
emissions to compare climate response across scenarios and provides Global and regional climate models are important sources of climate
a link to the emissions pathways assessment in WGIII. The advantage information at the regional scale. Since AR5, a more comprehensive
of using cumulative CO2 emissions is that it is an inherent emissions assessment of past and future evolution of a range of climate variables
scenario characteristic rather than an outcome of the scenario-based on a regional scale has been enabled by the increased availability
projections, where uncertainties in the cause–effect chain from of coordinated ensemble regional climate model projections and
emissions to temperature change are important (Figure TS.4), for improvements in the level of sophistication and resolution of
example, the uncertainty in ERF and TCR. Cumulative CO2 emissions global and regional climate models. This has been complemented
can also provide a link to the assessments of mitigation options. by observational, attribution and sectoral-vulnerability studies
Cumulative CO2 emissions do not carry information about non-CO2 informing, for instance, about impact-relevant tolerance thresholds.
emissions, although these can be included with specific emissions {10.3.3, 11.9, 12.1, 12.3, 12.6, Atlas.3–Atlas.11}
metrics to estimate CO2-equivalent emissions. (Section TS.3.3) {1.3.2,
1.6, 4.6.2, 5.5, 7.6} Multiple lines of evidence derived from observations, model simulations
and other approaches can be used to construct climate information
on a regional scale as described in detail in Sections TS.4.1.1 and
TS.1.4 From Global to Regional Climate Information for TS.4.1.2. Depending on the phenomena and specific context, these
Impact and Risk Assessment sources and methodologies include theoretical understanding of the
relevant processes, drivers and feedbacks of climate at regional scale; TS
The AR6 WGI Report has an expanded focus on regional trends in observed data from multiple datasets; and the attribution
information supported by the increased availability of of these trends to specific drivers. Furthermore, simulations from
coordinated regional climate model ensemble projections different model types (including global and regional climate models,
and improvements in the sophistication and resolution emulators, statistical downscaling methods, etc.) and experiments
of global and regional climate models (high confidence). (e.g., CMIP, CORDEX, and large ensembles of single-model simulations
Multiple lines of evidence can be used to construct climate with different initial conditions), attribution methodologies and other
information on a global to regional scale and can be further relevant local knowledge (e.g., indigenous knowledge) are utilized
distilled in a co-production process to meet user needs (high (see Box TS.11). {1.5.3, 1.5.4, Cross-Chapter Box 7.1, 10.2–10.6, 11.2,
confidence). To better support risk assessment, a common Atlas.1.4, Cross-Chapter Box 10.3}
risk framework across all three Working Groups has been
implemented in AR6, and low-likelihood but high-impact From the multiple lines of evidence, climate information can be
outcomes are explicitly addressed in WGI by using physical distilled in a co-production process that involves users, related
climate storylines (see Core Concepts Box). stakeholders and producers of climate information, considering the
specific context of the question at stake, the underlying values and
Climatic impact-drivers are physical climate system the challenge of communicating across different communities. The
conditions (e.g., means, events, extremes) that affect co-production process is an essential part of climate services, which
an element of society or ecosystems. They are the WGI are discussed in Section TS.4.1.2. {10.5, 12.6, Cross-Chapter Box 12.2}
contribution to the risk framing without anticipating
whether their impact provides potential opportunities With the aim of informing decision-making at local or regional
or is detrimental (i.e., as for hazards). Many global and scales, a common risk framework has been implemented in AR6.
regional climatic impact-drivers have a direct relation to Methodologies have been developed to construct more impact- and
global warming levels (high confidence). {1.4.4, 1.5.2–1.5.4, risk-relevant climate change information tailored to regions and
Cross-Chapter Box 1.3, 4.8, 10.1, 10.5.1, Box 10.2, Cross- stakeholders. Physical storyline approaches are used in order to
Chapter Box 10.3, 11.2.4, 11.9, Box 11.2, Cross-Chapter build climate information based on multiple lines of evidence, and
Box 11.1, 12.1–12.3, 12.6, Cross-Chapter Boxes 12.1 and which can explicitly address physically plausible, but low-likelihood,
12.2, Atlas.1.3.3–1.3.4, Atlas.1.4, Atlas.1.4.4} high-impact outcomes and uncertainties related to climate variability
for consideration in risk assessments (Figure TS.6). {Cross-Chapter
Climate change is a global phenomenon, but manifests differently Box 1.3, 4.8, Box 9.4, 10.5, Box 10.2, Box 11.2, 12.1–12.3, 12.6,
in different regions. The impacts of climate change are generally Glossary}
experienced at local, national and regional scales, and these are
also the scales at which decisions are typically made. Robust climate The climatic impact-driver framework developed in AR6 supports
change information is increasingly available at regional scales for an assessment of changing climate conditions that are relevant
impact and risk assessments. Depending on the climate information for sectoral impacts and risk assessment. Climatic impact-drivers
context, geographical regions in AR6 may refer to larger areas, such (CIDs) are physical climate system conditions (e.g., means, extremes,
57
Technical Summary
High ECS
high warming SSP2-4.5
SSP1-2.6
Equilibrium climate sensitivity (ECS, °c)
5
SSP1-1.9
Mid-range ECS
SSP3-7.0
Likely range SSP2-4.5
3
SSP1-2.6
SSP1-1.9
2 1980 2000 2020 2040 2060 2080
SSP5-8.5
Low-likelihood SSP3-7.0
1
Low ECS
Heat
warning
index
Changes in
extreme
rainfall
0 5 10 15 20 25 30
% more rain on wettest day of the year
Figure TS.6 | A graphical abstract for key aspects of the Technical Summary. The intent of this figure is to summarize many different aspects of the Technical Summary
related to observed and projected changes in global temperature and associated regional changes in climatic impact-drivers relevant for impact and risk assessment. Top left:
a schematic representation of the likelihood for equilibrium climate sensitivity (ECS), consistent with the AR6 assessment (see Chapter 7 and Section TS.3). ECS values above
5°C and below 2°C are termed low-likelihood, high warming (LLHW) and low-likelihood, low warming, respectively (Box TS.3). Top right: Observed (see Cross-Section Box
TS.1) and projected global surface temperature changes, shown as global warming levels (GWLs) relative to 1850–1900, using the assessed 95% (top), 50% (middle) and 5%
(bottom) likelihood time series (see Chapter 4 and Section TS.2). Bottom panels show maps of Coupled Model Intercomparison Project Phase 6 (CMIP6) median projections of
two climatic impact-drivers (CIDs, see Section TS.1.4) at three different GWLs (columns for 1.5, 2 and 4°C) for the AR6 land regions (see Chapters 1, 10, and Atlas and Section
TS.4). The heat warning index is the number of days per year averaged across each region at which a heat warning for human health at level ‘danger’ would be issued according
to the U.S. National Oceanic and Atmospheric Administration (NOAA) (NOAA HI41, see Chapter 12 and Annex VI). The maps of extreme rainfall changes show the percentage
change in the amount of rain falling on the wettest day of a year (Rx1day, relative to 1995–2014, see Chapter 11) averaged across each region when the respective GWL is
reached. Additional CIDs are discussed in Section TS.4. {1.4.4, Box 4.1, 7.5, 11.4.3, 12.4}
58
Technical Summary
events) that affect an element of society or ecosystems and are thus Many global- and regional-scale CIDs, including extremes, have
a potential priority for providing climate information. For instance, a direct relation to global warming levels (GWLs) and can thus inform
the heat index used by the U.S. National Oceanic and Atmospheric the hazard component of ‘Representative Key Risks’ and ‘Reasons for
Administration (NOAA HI) for issuing heat warnings is a CID index Concern’ assessed by AR6 WGII. These include heat, cold, wet and dry
that can be associated with adverse human health impacts due to hazards, both mean and extremes; cryospheric hazards (snow cover,
heat stress (see Figure TS.6). Depending on system tolerance, CIDs and ice extent, permafrost) and oceanic hazards (marine heatwaves)
their changes can be detrimental (i.e., hazards in the risk framing), (high confidence) (Figure TS.6). Establishing links between specific
beneficial, neutral, or a mixture of each across interacting system GWLs with tipping points and irreversible behaviour is challenging
elements, regions and sectors (aligning with WGII Sectoral Chapters due to model uncertainties and lack of observations, but their
2–8). Each sector is affected by multiple CIDs, and each CID affects occurrence cannot be excluded, and their likelihood of occurrence
multiple sectors. Climate change has already altered CID profiles and generally increases at greater warming levels (Box TS.1, Section TS.9).
resulted in shifting magnitude, frequency, duration, seasonality and {11.2.4, Box 11.2, Cross-Chapter Boxes 11.1 and 12.1}
spatial extent of associated indices (high confidence) (see regional
details in Section TS.4.3). {12.1–12.4, Table 12.1, Table 12.2, Annex VI}
This box synthesizes the outcomes of the assessment of past, current and future global surface temperature. Global mean surface
temperature (GMST) and global surface air temperature (GSAT) are the two primary metrics of global surface temperature used to
estimate global warming in IPCC reports. GMST merges sea surface temperature (SST) over the ocean and 2 m air temperature over
land and sea ice areas and is used in most paleo, historical and present-day observational estimates. The GSAT metric is 2 m air TS
temperature over all surfaces and is the diagnostic generally used from climate models. Changes in GMST and GSAT over time differ
by at most 10% in either direction (high confidence), but conflicting lines of evidence from models and direct observations, combined
with limitations in theoretical understanding, lead to low confidence in the sign of any difference in long-term trend. Therefore,
long-term changes in GMST/GSAT are presently assessed to be identical, with expanded uncertainty in GSAT estimates. Hence the term
global surface temperature is used in reference to both quantities in the text of the TS and SPM. {Cross-Chapter Box 2.3}
Global surface temperature has increased by 0.99 [0.84 to 1.10] °C from 1850–1900 to the first two decades of the
21st century (2001–2020) and by 1.09 [0.95 to 1.20] °C from 1850–1900 to 2011–2020. Temperatures as high as
during the most recent decade (2011–2020) exceed the warmest centennial-scale range reconstructed for the present
interglacial, around 6500 years ago [0.2°C to 1°C] (medium confidence). The next most recent warm period was
about 125,000 years ago during the last interglacial when the multi-centennial temperature range [0.5°C to 1.5°C]
encompasses the 2011–2020 values (medium confidence). The likely range of human-induced change in global surface
temperature in 2010–2019 relative to 1850–1900 is 0.8°C to 1.3°C, with a central estimate of 1.07°C, encompassing
the best estimate of observed warming for that period, which is 1.06°C with a very likely range of [0.88°C to 1.21°C],
while the likely range of the change attributable to natural forcing is only –0.1°C to +0.1°C.
Compared to 1850–1900, average global surface temperature over the period 2081–2100 is very likely to be higher by
[1.0°C to 1.8°C] in the low CO2 emissions scenario SSP1-1.9 and by [3.3°C to 5.7°C] in the high CO2 emissions scenario
SSP5-8.5. In all scenarios assessed here except SSP5-8.5, the central estimate of 20-year averaged global surface
warming crossing the 1.5°C level lies in the early 2030s, which is in the early part of the likely range (2030–2052)
assessed in SR1.5. It is more likely than not that under SSP1-1.9, global surface temperature relative to 1850–1900 will
remain below 1.6°C throughout the 21st century, implying a potential temporary overshoot of 1.5°C global warming
of no more than 0.1°C. Global surface temperature in any individual year could exceed 1.5°C relative to 1850–1900
by 2030 with a likelihood between 40% and 60% across the scenarios considered here (medium confidence). A 2°C
increase in global surface temperature relative to 1850–1900 will be crossed under SSP5-8.5 but is extremely unlikely
to be crossed under SSP1-1.9. Periods of reduced and increased global surface temperature trends at decadal time
scales will continue to occur in the 21st century (very high confidence). The effect of strong mitigation on 20-year
global surface temperature trends would be likely to emerge during the near term (2021–2040), assuming no major
volcanic eruptions occur. (Figure TS.8, Cross-Section Box TS.1, Figure 1) {2.3, 3.3, 4.3, 4.4, 4.5, 4.6, 7.3}
59
Technical Summary
to 1986–2005 is estimated at 0.08 [–0.01 to 0.12] °C. Global surface temperature increased from 1850–1900 to 1995–2014 by 0.85
[0.69 to 0.95] °C, between 1850–1900 and the first two decades of the 21st century (2001–2020) by 0.99 [0.84 to 1.20] °C, and to
the most recent decade (2011–2020) by 1.09 [0.95 to 1.20] °C. Each of the last four decades has in turn been warmer than any decade
that preceded it since 1850. Temperatures have increased faster over land than over the ocean since 1850–1900, with warming to
2011–2020 of 1.59 [1.34 to 1.83] °C over land and 0.88 [0.68 to 1.01] °C over the ocean. {2.3.1, Cross-Chapter Box 2.3}
Global surface temperature during the period 1850–1900 is used as an approximation for pre-industrial conditions for consistency
with AR5 and AR6 Special Reports, whilst recognizing that radiative forcings have a baseline of 1750 for the start of anthropogenic
influences. It is likely that there was a net anthropogenic forcing of 0.0–0.3 Wm–2 in 1850–1900 relative to 1750 (medium confidence),
and from the period around 1750 to 1850–1900, there was a change in global surface temperature of around 0.1°C (likely range
–0.1 to +0.3°C, medium confidence), with an anthropogenic component of 0.0°C to 0.2°C (likely range, medium confidence). {Cross-
Chapter Box 1.2, 7.3.5}
Global surface temperature has evolved over geological time (Figure TS.1, Box TS.2). Beginning approximately 6500 years ago, global
surface temperature generally decreased, culminating in the coldest multi-century interval of the post-glacial period (since roughly
7000 years ago), which occurred between around 1450 and 1850 (high confidence). Over the last 50 years, global surface temperature
has increased at an observed rate unprecedented in at least the last two thousand years (high confidence). Temperatures as high as
during the most recent decade (2011–2020) exceed the warmest centennial-scale range reconstructed for the present interglacial,
around 6500 years ago [0.2°C to 1°C] (medium confidence). The next most recent warm period was about 125,000 years ago during
TS the Last Interglacial when the multi-centennial temperature range [0.5°C to 1.5°C] encompasses the 2011–2020 values (medium
confidence) (Cross-Section Box TS.1, Figure 1). During the mid-Pliocene Warm Period, around 3.3–3.0 million years ago, global surface
temperature was 2.5°C to 4°C warmer (medium confidence). {2.3.1, Cross-Chapter Box 2.1 and 2.4}
Current Warming
There is very high confidence that the CMIP6 model ensemble reproduces observed global surface temperature trends and variability
since 1850 with errors small enough to allow for detection and attribution of human-induced warming. The CMIP6 multi-model mean
global surface warming between 1850–1900 and 2010–2019 is close to the best estimate of observed warming, though some CMIP6
models simulate a warming that is outside the assessed very likely observed range. {3.3.1}
The likely range of human-induced change in global surface temperature in 2010–2019 relative to 1850–1900 is 0.8°C to 1.3°C,
with a central estimate of 1.07°C (Figure Cross-Section Box TS.1, Figure 1), encompassing the best estimate of observed warming
for that period, which is 1.06°C with a very likely range of [0.88°C to 1.21°C], while the likely range of the change attributable
to natural forcing is only –0.1°C to +0.1°C. This assessment is consistent with an estimate of the human-induced global surface
temperature rise based on assessed ranges of perturbations to the top of the atmosphere (effective radiative forcing) and with metrics
of feedbacks of the climate response (equilibrium climate sensitivity and the transient climate response). Over the same period, well-
mixed greenhouse gas forcing likely warmed global surface temperature by 1.0°C to 2.0°C, while aerosols and other anthropogenic
forcings likely cooled global surface temperature by 0.0°C to 0.8°C. {2.3.1, 3.3.1, 7.3.5, Cross-Chapter Box 7.1}
The observed slower increase in global surface temperature (relative to preceding and following periods) in the 1998–2012 period,
sometimes referred to as ‘the hiatus’, was temporary (very high confidence). The increase in global surface temperature during the
1998–2012 period is also greater in the data sets used in the AR6 assessment than in those available at the time of AR5. Using these
updated observational data sets and a like-for-like consistent comparison of simulated and observed global surface temperature, all
observed estimates of the 1998–2012 trend lie within the very likely range of CMIP6 trends. Furthermore, the heating of the climate
system continued during this period, as reflected in the continued warming of the global ocean (very high confidence) and in the
continued rise of hot extremes over land (medium confidence). Since 2012, global surface temperature has risen strongly, with the
past five years (2016–2020) being the hottest five-year period between 1850 and 2020 (high confidence). {2.3.1, 3.3.1, 3.5.1, Cross-
Chapter Box 3.1}
60
Technical Summary
0.5
Mid-Holocene
0.0
Last Interglacial
Latest decade
−0.5
−10,000 −6000 −2000 1000 1400 1800 1900 2000
Year (BCE) Year (CE)
Age: Agricultural Historical Industrial
Resolution: Centuries Decades Annual
(b) Observed and projected warming are stronger over (c) Global surface temperature has risen more than 1°C
land than oceans, and strongest in the Arctic from 1850–1900
1.0 HadCRUT .5.0 NOAAGlobalTemp
Kadow et al. Berkeley Earth
1981–2020 0.5
TS
Global surface temperature relative to 1850–1900 (°C)
0.0
−0.5
1850 1900 1950 2000
(d) Internal variability will influence near-term warming rates
2.5 CMIP6 historical
HadCRUT5
assessed SSP1-2.6
2.0 assessed SSP2-4.5
assessed SSP3-7.0
1.5 models 1-4
−0.6 −0.4 −0.2 −0.1 0.0 0.1 0.2 0.4 0.6 1.0
SSP1-1.9
SSP1-2.6
SSP2-4.5 5
4 SSP3-7.0
Relative to 1850–1900 (°C)
SSP5-8.5
4
3
3
2
1 2
−6 −4 −3 −2 −1 −0.5 0 0.5 1 2 3 4 6 1.5 °C
Total change (°C)
1
0
2000–2019 2020–2039 2040–2059 2060–2079 2081–2100
Cross-Section Box TS.1, Figure 1 | Earth’s surface temperature history and future with key findings annotated within each panel.
61
Technical Summary
During the near term (2021–2040), a 1.5°C increase in global surface temperature, relative to 1850–1900, is very likely to occur
in scenario SSP5-8.5, likely to occur in scenarios SSP2-4.5 and SSP3-7.0, and more likely than not to occur in scenarios SSP1-1.9 and
SSP1-2.6. The time of crossing a warming level is defined here as the midpoint of the first 20-year period during which the average
global surface temperature exceeds the level. In all scenarios assessed here except SSP5-8.5, the central estimate of crossing the 1.5°C
TS level lies in the early 2030s. This is in the early part of the likely range (2030–2052) assessed in SR1.5, which assumed continuation of
the then-current warming rate; this rate has been confirmed in the AR6. Roughly half of this difference arises from a larger historical
warming diagnosed in AR6. The other half arises because for central estimates of climate sensitivity, most scenarios show stronger
warming over the near term than was estimated as ‘current’ in SR1.5 (medium confidence). When considering scenarios similar to SSP1-
1.9 instead of linear extrapolation, the SR1.5 estimate of when 1.5°C global warming is crossed is close to the central estimate reported
here. (Cross-Section Box TS.1, Table 1) {2.3.1, Cross-Chapter Box 2.3, 3.3.1, 4.3.4, Box 4.1}
It is more likely than not that under SSP1-1.9, global surface temperature relative to 1850–1900 will remain below 1.6°C throughout
the 21st century, implying a potential temporary overshoot of 1.5°C global warming of no more than 0.1°C. If climate sensitivity lies
near the lower end of the assessed very likely range, crossing the 1.5°C warming level is avoided in scenarios SSP1-1.9 and SSP1-2.6
(medium confidence). Global surface temperature in any individual year, in contrast to the 20-year average, could by 2030 exceed
1.5°C relative to 1850–1900 with a likelihood between 40% and 60%, across the scenarios considered here (medium confidence).
(Cross-Section Box TS.1, Table 1) {4.3.4, 4.4.1, Box 4.1, 7.5}
During the 21st century, a 2°C increase in global surface temperature relative to 1850–1900 will be crossed under SSP5-8.5 and
SSP3-7.0, is extremely likely to be crossed under SSP2-4.5, but is unlikely to be crossed under SSP1-2.6 and extremely unlikely to be
crossed under SSP1-1.9. For the mid-term period 2041–2060, this 2°C global warming level is very likely to be crossed under SSP5-8.5,
likely to be crossed under SSP3-7.0, and more likely than not to be crossed under SSP2-4.5. (Cross-Section Box TS.1, Table 1) {4.3.4}
Events of reduced and increased global surface temperature trends at decadal time scales will continue to occur in the 21st century
but will not affect the centennial-scale warming (very high confidence). If strong mitigation is applied from 2020 onward as reflected
in SSP1-1.9, its effect on 20-year trends in global surface temperature would likely emerge during the near term (2021–2040),
measured against an assumed non-mitigation scenario such as SSP3-7.0 or SSP5-8.5. All statements about crossing the 1.5°C level
assume that no major volcanic eruption occurs during the near term (Cross-Section Box TS.1, Table 1). {2.3.1, Cross-Chapter Box 2.3,
4.3.4, 4.4.1, 4.6.3, Box 4.1}
Compared to 1850–1900, average global surface temperature over the period 2081–2100 is very likely to be higher by [1.0°C to
1.8°C] in the low CO2 emissions scenario SSP1-1.9 and by [3.3°C to 5.7°C] in the high CO2 emissions scenario SSP5-8.5. For the
scenarios SSP1-2.6, SSP2-4.5, and SSP3-7.0, the corresponding very likely ranges are [1.3°C to 2.4°C], [2.1°C to 3.5°C], and [2.8°C
to 4.6°C], respectively. The uncertainty ranges for the period 2081–2100 continue to be dominated by the uncertainty in equilibrium
climate sensitivity and transient climate response (very high confidence) (Cross-Section Box TS.1, Table 1). {4.3.1, 4.3.4, 4.4.1, 7.5}
The CMIP6 models project a wider range of global surface temperature change than the assessed range (high confidence); furthermore,
the CMIP6 global surface temperature increase tends to be larger than that in CMIP5 (very high confidence). {4.3.1, 4.3.4, 4.6.2, 7.5.6}
62
Technical Summary
Cross-Section Box TS.1, Table 1 | Assessment results for 20-year averaged change in global surface temperature based on multiple lines of
evidence. The change is displayed in °C relative to the 1850–1900 reference period for selected time periods (first three rows), and as the first 20-year period during
which the average global surface temperature change exceeds the specified level relative to the period 1850–1900 (last four rows). The entries give both the central
estimate and, in parentheses, the very likely (5–95%) range. An entry n.c. means that the global warming level is not crossed during the period 2021–2100.
Mid-term,
1.6 [1.2 to 2.0] 1.7 [1.3 to 2.2] 2.0 [1.6 to 2.5] 2.1 [1.7 to 2.6] 2.4 [1.9 to 3.0]
2041–2060
Long term,
1.4 [1.0 to 1.8] 1.8 [1.3 to 2.4] 2.7 [2.1 to 3.5] 3.6 [2.8 to 4.6] 4.4 [3.3 to 5.7]
2081–2100
TS
TS.2 Large-scale Climate Change: Mean substantial reductions in global GHG emissions. Continued
Climate, Variability and Extremes GHG emissions greatly increase the likelihood of potentially
irreversible changes in the global climate system (Box TS.9),
This section summarizes knowledge about observed and projected in particular with respect to the contribution of ice sheets
large-scale climate change (including variability and extremes), to global sea level change (high confidence). {2.3, 3.8, 4.3,
drivers and attribution of observed changes to human activities. It 4.6, 4.7, 7.2–7.4, Cross-Chapter Box 7.1, 9.2–9.6}
describes observed and projected large-scale changes associated
with major components of the climate system: atmosphere, ocean Earth system model simulations of the historical period since 1850
(including sea level change), land, biosphere and cryosphere, and the are only able to reproduce the observed changes in key climate
carbon, energy and water cycles. In each subsection, reconstructed indicators when anthropogenic forcings are included (Figure TS.7).
past changes, observed and attributed recent changes, and projected Taken together with numerous formal attribution studies across an
near- and long-term changes to mean climate, variability and even broader range of indicators and theoretical understanding,
extremes are presented, where possible, in an integrated way. See this underpins the unequivocal attribution of observed warming of
Section TS.1.3.1 for information on the scenarios used for projections. the atmosphere, ocean, and land to human influence (Table TS.1).
{2.3, 3.8}
63
Technical Summary
Global
Near-surface air temperature
over land Near-surface air temperature Ocean heat content
Africa Arctic
60°N–90°N
Figure TS.7 | Simulated and observed changes compared to the 1850–1900 average in key large-scale indicators of climate change across the climate
system, for continents, ocean basins and globally up to 2014. The intent of this figure is to compare the observed and simulated changes over the historical period
for a range of variables and regions, with and without anthropogenic forcings, for attribution. Black lines show observations, orange lines and shading show the multi-model
mean and 5–95th percentile ranges for Coupled Model Intercomparison Project Phase 6 (CMIP6) historical simulations including anthropogenic and natural forcing, and green
lines and shading show corresponding ensemble means and 5–95th percentile ranges for CMIP6 natural-only simulations. Observations after 2014 (including, for example, a
strong subsequent decrease of Antarctic sea ice area that leads to no significant overall trend since 1979) are not shown because the CMIP6 historical simulations end in 2014.
A 3-year running mean smoothing has been applied to all observational time series. {3.8, Figure 3.41}
64
Technical Summary
Table TS.1 | Assessment of observed changes in large-scale indicators of mean climate across climate system components and their attribution to human
influence. The colour coding indicates the assessed confidence in/likelihood of the human contribution as a driver or main driver19 (main driver is specified in that case) where
available (see colour key). Otherwise, explanatory text is provided in cells with white background. The relevant chapter section with more detailed information is listed in each
table cell.
Warming of global mean surface air temperature since Likely range of human contribution (0.8°C–1.3°C) encom-
{2.3.1, Cross-Chapter Box 2.3}
1850–1900 passes observed warming (0.9°C–1.2°C) {3.3.1}
See text description medium confidence likely/high confidence very likely extremely likely virtually certain fact
Future climate change across a range of atmospheric, cryospheric, centuries to millennia. Furthermore, it is likely that at least one large
oceanic and biospheric indicators depends upon future emissions volcanic eruption will occur during the 21st century. Such an eruption
pathways. Outcomes for a broad range of indicators increasingly would reduce global surface temperature for several years, decrease
diverge through the 21st century across the different SSPs (Section land precipitation, alter monsoon circulation and modify extreme
TS.1.3.1, Figure TS.8). Due to the slow response of the deep ocean precipitation, at both global and regional scales. {4.3, 4.7, 9.4, 9.6,
and ice sheets, this divergence continues long after 2100, and 21st Cross-Chapter Box 4.1}
century emissions choices will have implications for GMSL rise for
19 Throughout this Technical Summary, ‘main driver’ means responsible for more than 50% of the change.
65
Technical Summary
Recent and future change of four key indicators of the climate system
Atmospheric temperature, ocean heat content, Arctic summer sea ice, and land precipitation
(a) Global surface air temperature (b) Global ocean heat content and thermosteric sea level
4
CMIP6 Emulator
Future (assessed) 3
3 Future (assessed) 0.3
2 2 0.2
m
°C
YJ
1 1 Past (observed) 0.1
Past (observed)
0 Past (simulated) 0 Past (simulated) 0
–1 –1 -0.1
1950 2000 [2000/ [2040/ [2080/ Change 1950 2000 2050 2100 Change
2019] 2059] 2100] in 2100 in 2100
(c) Arctic September sea ice area (d) Global land precipitation
10 20
8 Past (simulated) Future (CMIP6)
Past (observed)
10
6
million km2
Past (observed)
Future (CMIP6)
%
4 0
2
Practically sea ice free Past (simulated)
0 –10
1950 2000 2050 2100 Change 1950 2000 2050 2100 Change
in 2100
in 2100
Figure TS.8 | Observed, simulated and projected changes compared to the 1995–2014 average in four key indicators of the climate system through to
2100 differentiated by Shared Socio-economic Pathway (SSP) scenario. The intent of this figure is to show how future emissions choices impact key, iconic large-scale
indicators and to highlight that our collective choices matter. Past simulations are based on the Coupled Model Intercomparison Project Phase 6 (CMIP6) multi-model ensemble.
Future projections are based on the assessed ranges based upon multiple lines of evidence for (a) global surface temperature (Cross-Section Box TS.1) and (b) global ocean heat
content and the associated thermosteric sea level contribution to global mean sea level change (right-hand axis) using a climate model emulator (Cross-Chapter Box 7.1), and
CMIP6 simulations for (c) Arctic September sea ice and (d) global land precipitation. Projections for SSP1-1.9 and SSP1-2.6 show that reduced greenhouse gas emissions lead
to a stabilization of global surface temperature, Arctic sea ice area and global land precipitation over the 21st century. Projections for SSP1-2.6 show that emissions reductions
have the potential to substantially reduce the increase in ocean heat content and thermosteric sea level rise over the 21st century but that some increase is unavoidable. The
brackets in the x axis in panel (a) indicate assessed 20-year-mean periods. {4.3, Figure 4.2, 9.3, 9.6, Figure 9.6}
Observational records show changes in a wide range of climate human influence is the main contributor to the observed increase
extremes that have been linked to human influence on the climate (decrease) in the likelihood and severity of hot (cold) extremes (Table
system (Table TS.2). In many cases, the frequency and intensity of TS.2). The frequency of extreme temperature and precipitation events
future changes in extremes can be directly linked to the magnitude in the current climate will change with warming, with warm extremes
of future projected warming. Changes in extremes have been becoming more frequent (virtually certain), cold extremes becoming
widespread over land since the 1950s, including a virtually certain less frequent (extremely likely) and precipitation extremes becoming
global increase in extreme air temperatures and a likely intensification more frequent in most locations (very likely). {9.6.4, 11.2, 11.3, 11.4,
in global-scale extreme precipitation. It is extremely likely that 11.6, 11.7, 11.8, 11.9, Box 9.2}
66
Technical Summary
Table TS.2 | Summary table on observed changes in extremes, their attribution since 1950 (except where stated otherwise), and projected changes at
+1.5°C, +2°C and +4°C of global warming, on global and continental scales. An increase in warm/hot extremes refers to warmer and/or more frequent hot days and
nights and warm spells/heatwaves, over most land areas. A decrease in cold extremes refers to warmer and/or fewer cold days and nights and cold spells/cold waves, over most
land areas. Drought events are relative to a predominant fraction of land area. For tropical cyclones, observed changes and attribution refer to Categories 3–5, while projected
changes refer to Categories 4–5. Tables 11.1 and 11.2 are more detailed versions of this table, containing, in particular, information on regional scales. In general, higher
warming levels also imply stronger projected changes for indicators where the confidence level does not depend on the warming level and the table does not explicitly quantify
the global sensitivity. See also Box TS.10. {9.6, Box 9.2, 11.3, 11.7}
↑ ↑ ↑
Heavy precipitation events: Frequency, Over majority of land
Main driver of the
observed intensification
↑
intensity and/or amount regions with good in most land regions in most land regions
of heavy precipitation in
observational coverage
land regions
↑ ↑ ↑
Agricultural and ecological droughts: ↑ in more regions in more regions in more regions
Intensity and/or frequency in some regions in some regions compared to observed compared to 1.5°C compared to 2°C of
changes of global warming global warming
medium confidence likely/high confidence very likely extremely likely virtually certain
TS.2.2 Changes in the Drivers of the Climate System The total anthropogenic effective radiative forcing (ERF)
in 2019, relative to 1750, was 2.72 [1.96 to 3.48] W m–2
Since 1750, changes in the drivers of the climate system (medium confidence) and has likely been growing at an
are dominated by the warming influence of increases in increasing rate since the 1970s. {2.2, 6.4, 7.2, 7.3}
atmospheric GHG concentrations and a cooling influence
from aerosols, both resulting from human activities. In Solar activity since 1900 was high but not exceptional compared to
comparison there has been negligible long-term influence the past 9000 years (high confidence). The average magnitude and
from solar activity and volcanoes. Concentrations of CO2, variability of volcanic aerosols since 1900 has not been unusual
methane (CH4), and nitrous oxide (N2O) have increased to compared to at least the past 2500 years (medium confidence).
levels unprecedented in at least 800,000 years, and there However, sporadic strong volcanic eruptions can lead to temporary
is high confidence that current CO2 concentrations have not drops in global surface temperature lasting 2–5 years. {2.2.1, 2.2.2,
been experienced for at least 2 million years. Global mean 2.2.8, Cross-Chapter Box 4.1}
concentrations of anthropogenic aerosols peaked in the
late 20th century and have slowly declined since in northern Atmospheric CO2 concentrations have changed substantially
mid-latitudes, although they continue to increase in South over millions of years (Figure TS.1). Current levels of atmospheric
Asia and East Africa (high confidence). CO2 have not been experienced for at least 2 million years (high
67
Technical Summary
(a) Last time CO2 levels were as high as present was at least 2 million years ago
δ11B-foraminifera
δ13C-alkenone
450
Ant. ice core
409.9
CO2 (ppm) 350
250
150
3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0
Age (millions of years ago)
380 NOAA
CO2 (ppm)
360 EDML
360 CSIRO
340 SIO
340
320
300 1866.3 320
1800 300 1866.3
TS 280 1800
1600 280
1600
CH4 (ppb)
1400
CH4 (ppb)
GISP2 1400
1200
WAIS Divide 1200
1000
Law Dome NOAA 1000
800 CSIRO
AGAGE 800
600 UCI
340 600
332.1 340
320 332.1
320
N2O (ppb)
GISP2
N2O (ppb)
0
W m−2
−2
(W m−2 per decade−1)
0.5
0.4
Rate of change anthropogenic ERF 0.3
−4 0.2
0.1
0.0
Figure TS.9 | Changes in well-mixed greenhouse gas (WMGHG) concentrations and effective radiative forcing (EFR). The intent of this figure is to show that the
changes of the main drivers of climate system over the industrial period are exceptional in a long-term context. (a) Changes in carbon dioxide (CO2) from proxy records over the
past 3.5 million years. (b) Changes in all three WMGHGs from ice core records over the Common Era. (c) Directly observed WMGHG changes since the mid-20th century. (d)
Evolution of ERF and components since 1750. Further details on data sources and processing are available in the associated FAIR data table. {2.2, Figures 2.3, 2.4 and 2.10}
68
Technical Summary
confidence, Figure TS.9a). Over 1750–2019, CO2 increased by 131.6 the late 20th century, with low confidence in the magnitude of
± 2.9 ppm (47.3%). The centennial rate of change of CO2 since 1850 post-2014 changes due to conflicting evidence (Section TS.3.1).
has no precedent in at least the past 800,000 years (Figure TS.9), {2.2.6, 6.2.1, 6.3.5, 6.4.1, 7.3.3}
and the fastest rates of change over the last 56 million years were
at least a factor of four lower (low confidence) than over 1900– There is high confidence that tropospheric ozone has been increasing
2019. Several networks of high-accuracy surface observations from 1750 in response to anthropogenic changes in ozone precursor
show that concentrations of CO2 have exceeded 400 ppm, reaching emissions (nitrogen oxides, carbon monoxide, non-methane volatile
409.9 (± 0.3) ppm in 2019 (Figure TS.9c). The ERF from CO2 in 2019 organic compounds, and methane), but with medium confidence in
(relative to 1750) was 2.16 Wm–2. {2.2.3, 5.1.2, 5.2.1, 7.3} the magnitude of this change, due to limited observational evidence
and knowledge gaps. Since the mid-20th century, tropospheric
By 2019, concentrations of CH4 reached 1866.3 (± 3.3) ppb ozone surface concentrations have increased by 30–70% across the
(Figure TS.9c). The increase since 1750 of 1137 ± 10 ppb (157.8%) far Northern Hemisphere (medium confidence); since the mid-1990s,
exceeds the range over multiple glacial–interglacial transitions of the free tropospheric ozone has increased by 2–7% per decade in most
past 800,000 years (high confidence). In the 1990s, CH4 concentrations northern mid-latitude regions and 2–12% per decade in sampled
plateaued, but started to increase again around 2007 at an average tropical regions. Future changes in surface ozone concentrations will
rate of 7.6 ± 2.7 ppb yr –1 (2010–2019; high confidence). There is high be primarily driven by changes in precursor emissions rather than
confidence that this recent growth is largely driven by emissions climate change (high confidence). Stratospheric ozone has declined
from fossil fuel exploitation, livestock, and waste, with ENSO driving between 60°S–60°N by 2.2% from 1964–1980 to 2014–2017
multi-annual variability of wetland and biomass burning emissions. (high confidence), with the largest declines during 1980–1995.
In 2019, ERF from CH4 was 0.54 Wm–2. {2.2.3, 5.2.2, 7.3} The strongest loss of stratospheric ozone continues to occur in
austral spring over Antarctica (ozone hole), with emergent signs of
Since 1750, N2O increased by 62.0 ± 6.0 ppb, reaching a level of recovery after 2000. The 1750–2019 ERF for total (stratospheric and
332.1 (± 0.4) ppb in 2019. The increase since 1750 is of comparable tropospheric) ozone is 0.47 [0.24 to 0.71] W m−2, which is dominated TS
magnitude to glacial–interglacial fluctuations of the past by tropospheric ozone changes. {2.2.5, 6.3.2, 7.3.2, 7.3.5}
800,000 years (Figure TS.9c). N2O concentration trends since 1980
are largely driven by a 30% increase in emissions from the expansion The global mean abundance of hydroxyl (OH) radical, or ‘oxidizing
and intensification of global agriculture (high confidence). By 2019 capacity’, chemically regulates the lifetimes of many SLCFs, and
its ERF was 0.21 W m–2. {2.2.3, 5.2.3} therefore the radiative forcing of CH4, ozone, secondary aerosols
and many halogenated species. Model estimates suggest no
Halogenated gases consist of chlorofluorocarbons (CFCs), significant change in oxidizing capacity from 1850 to 1980 (low
hydrochlorofluorocarbons (HCFCs), hydrofluorocarbons (HFCs) and confidence). Increases of about 9% over 1980–2014 computed by
other gases, many of which can deplete stratospheric ozone and ESMs and carbon cycle models are not confirmed by observationally
warm the atmosphere. In response to controls on production and constrained inverse models, rendering an overall medium confidence
consumption mandated by the Montreal Protocol on Substances in stable OH or positive trends since the 1980s, and implying that OH
that Deplete the Ozone Layer and its amendments, the atmospheric is not the primary driver of recent observed growth in CH4. {6.3.6,
abundances of most CFCs have continued to decline since AR5. Cross-Chapter Box 5.2}
Abundances of HFCs, which are replacements for CFCs and HCFCs,
are increasing (high confidence), though increases of the major HCFCs Land use and land-cover change exert biophysical and biogeochemical
have slowed in recent years. The ERF from halogenated components effects. There is medium confidence that the biophysical effects
in 2019 was 0.41 Wm–2. {2.2.4, 6.3.4, 7.3.2} of land-use change since 1750, most notably the increase in
global albedo, have had an overall cooling on climate, whereas
Tropospheric aerosols mainly act to cool the climate system, biogeochemical effects (i.e., changes in GHG and volatile organic
directly by reflecting solar radiation, and indirectly by enhancing compound emissions or sinks) led to net warming. Overall land-use
cloud reflectance. Ice cores show increases in aerosols across the and land-cover ERF is estimated at –0.2 [–0.3 to –0.1] W m−2. {2.2.7,
Northern Hemisphere mid-latitudes since 1700 and reductions 7.3.4, SRCCL Section 2.5}
since the late 20th century (high confidence). Aerosol optical depth
(AOD), derived from satellite- and ground-based radiometers, The total anthropogenic ERF in 2019 relative to 1750 was 2.72 [1.96
has decreased since 2000 over the mid-latitude continents of to 3.48] W m−2 (Figure TS.9), dominated by GHGs (positive ERF) and
both hemispheres, but increased over South Asia and East Africa partially offset by aerosols (negative ERF). The rate of change of ERF
(high confidence). Trends in AOD are more pronounced from sub- likely has increased since the 1970s, mainly due to growing CO2
micrometre aerosols for which the anthropogenic contribution is concentrations and less negative aerosol ERF (Section TS.3.1). {2.2.8,
particularly large. Global carbonaceous aerosol budgets and trends 7.3}
remain poorly characterized due to limited observations, but black
carbon (BC), a warming aerosol component, is declining in several
regions of the Northern Hemisphere (low confidence). Total aerosol
ERF in 2019, relative to 1750, is −1.1 [−1.7 to −0.4] W m−2 (medium
confidence) and more likely than not became less negative since
69
Technical Summary
TS.2.3 Upper Air Temperatures and Atmospheric track in the Southern Hemisphere by 2100 under the high
Circulation CO2 emissions scenarios. It is likely that the proportion of
intense tropical cyclones has increased over the last four
The effects of human-induced climate change have decades and that this cannot be explained entirely by
been clearly identified in observations of atmospheric natural variability. There is low confidence in observed
temperature and some aspects of atmospheric circulation, recent changes in the total number of extratropical cyclones
and these effects are likely to intensify in the future. over both hemispheres. The proportion of tropical cyclones
Tropospheric warming and stratospheric cooling are that are intense is expected to increase (high confidence),
virtually certain to continue with continued net emissions but the total global number of tropical cyclones is expected
of greenhouse gases. Several aspects of the atmospheric to decrease or remain unchanged (medium confidence).
circulation have likely changed since the mid-20th century, {2.3, 3.3, 4.3, 4.4, 4.5, 8.3, 8.4, 11.7}
and human influence has likely contributed to the observed
poleward expansion of the Southern Hemisphere Hadley Cell The troposphere has warmed since at least the 1950s, and it is virtually
and very likely contributed to the observed poleward shift certain that the stratosphere has cooled. It is very likely that human-
of the Southern Hemisphere extratropical jet in summer. It induced increases in GHGs were the main driver of tropospheric
is likely that the mid-latitude jet will shift poleward and warming since 1979. It is extremely likely that anthropogenic
strengthen, accompanied by a strengthening of the storm forcing, both from increases in GHG concentrations and depletion of
Pressure (hPa)
Pressure (hPa)
Pressure (hPa)
ERA5 1979–2019 (DJF) Multi-model projections (DJF, SSP1-2.6) Multi-model projections (DJF, SSP3-7.0)
34 31
Pressure (hPa)
Pressure (hPa)
Pressure (hPa)
Figure TS.10 | Observed and projected upper air temperature and circulation changes. The intent of this figure is to visualize upper air temperature and circulation
changes and the similarity between observed and projected changes. Upper panels: (Left) Zonal cross-section of temperature trends for 2002–2019 in the upper troposphere
region for the ROM SAF radio-occultation dataset. (Middle) Change in the annual and zonal mean atmospheric temperature (°C) in 2081–2100 in SSP1-2.6 relative to
1995–2014 for 36 Coupled Model Intercomparison Project Phase 6 (CMIP6) models. (Right) the same in SSP3-7.0 for 32 models. Lower panels: (Left) Long-term mean (thin
black colour) and linear trend (colour) of zonal mean December–January–February (DJF) zonal winds for ERA5. (Middle) multi-model mean change in annual and zonal mean
wind (m s–1) in 2081–2100 in SSP1-2.6 relative to 1995–2014 based on 34 CMIP6 models. The 1995–2014 climatology is shown in contours with spacing of 10 m s–1. (Right)
the same for SSP3-7.0 for 31 models. {2.3.1; Figures 2.12 and 2.18; 4.5.1; Figure 4.2.6}
70
Technical Summary
stratospheric ozone due to ozone-depleting substances, was the main the Southern Hemisphere mid-latitude jet is likely to shift poleward
driver of upper stratospheric cooling since 1979. It is very likely that and strengthen under the SSP5-8.5 scenario relative to 1995–2014,
global mean stratospheric cooling will be larger for scenarios with accompanied by an increase in the SAM (Section TS.4.2.2). It is
higher atmospheric CO2 concentrations. In the tropics, since at least likely that wind speeds associated with extratropical cyclones will
2001 (when new techniques permit more robust quantification), the strengthen in the Southern Hemisphere storm track for SSP5-8.5.
upper troposphere has warmed faster than the near-surface (medium There is low confidence in the potential role of Arctic warming and
confidence) (Figure TS.10). There is medium confidence that most sea ice loss on historical or projected mid-latitude atmospheric
CMIP5 and CMIP6 models overestimate the observed warming in variability. {2.3.1, 3.3.3, 3.7.2, 4.3.3, 4.4.3, 4.5.1, 4.5.3, 8.2.2, 8.3.2,
the upper tropical troposphere over the period 1979–2014, in part Cross-Chapter Box 10.1}
because they overestimate tropical SST warming. It is likely that
future tropical upper tropospheric warming will be larger than at the It is likely that the proportion of major (Category 3–5) tropical
tropical surface. {2.3.1, 3.3.1, 4.5.1} cyclones (TCs) and the frequency of rapid TC intensification events
have increased over the past four decades. The average location
The Hadley Circulation has likely widened since at least the 1980s, of peak TC wind-intensity has very likely migrated poleward in
predominantly in the Northern Hemisphere, although there is only the western North Pacific Ocean since the 1940s, and TC forward
medium confidence in the extent of the changes. This has been translation speed has likely slowed over the contiguous USA since
accompanied by a strengthening of the Hadley Circulation in the 1900. It is likely that the poleward migration of TCs in the western
Northern Hemisphere (medium confidence). It is likely that human North Pacific and the global increase in TC intensity rates cannot be
influence has contributed to the poleward expansion of the zonal explained entirely by natural variability. There is high confidence
mean Hadley cell in the Southern Hemisphere since the 1980s, which that average peak TC wind speeds and the proportion of Category
is projected to further expand with global warming (high confidence). 4–5 TCs will increase with warming and that peak winds of the
There is medium confidence that the observed poleward expansion most intense TCs will increase. There is medium confidence that the
in the Northern Hemisphere is within the range of internal variability. average location where TCs reach their maximum wind-intensity will TS
{2.3.1, 3.3.3, 8.4.3} migrate poleward in the western North Pacific Ocean, while the total
global frequency of TC formation will decrease or remain unchanged
Since the 1970s, near-surface average winds have likely weakened with increasing global warming. {11.7.1}
over land. Over the ocean, near-surface average winds likely
strengthened over 1980–2000, but divergent estimates lead to There is low confidence in observed recent changes in the total
low confidence thereafter. Extratropical storm tracks have likely number of extratropical cyclones over both hemispheres. There is also
shifted poleward since the 1980s. There is low confidence in low confidence in past-century trends in the number and intensity of
projected poleward shifts of the Northern Hemisphere mid-latitude the strongest extratropical cyclones over the Northern Hemisphere
jet and storm tracks due to large internal variability and structural due to the large interannual-to-decadal variability and temporal and
uncertainty in model simulations. There is medium confidence in spatial heterogeneities in the volume and type of assimilated data
a projected decrease in the frequency of atmospheric blocking over in atmospheric reanalyses, particularly before the satellite era. Over
Greenland and the North Pacific in boreal winter in 2081–2100 under the Southern Hemisphere, it is likely that the number of extratropical
the SSP3-7.0 and SSP5-8.5 scenarios. There is high confidence that cyclones with low central pressures (<980 hPa) has increased since
Southern Hemisphere storm tracks and associated precipitation have 1979. The frequency of intense extratropical cyclones is projected to
migrated polewards over recent decades, especially in the austral decrease (medium confidence). Projected changes in the intensity
summer and autumn, associated with a trend towards more positive depend on the resolution of climate models (medium confidence).
phases of the Southern Annular Mode (SAM) (Section TS.4.2.2) and There is medium confidence that wind speeds associated with
the strengthening and southward shift of the Southern Hemisphere extratropical cyclones will change following changes in the storm
extratropical jet in austral summer. In the long term (2081–2100), tracks. {2.3.1, 3.3.3, 4.5.1, 4.5.3, 8.3.2, 8.4.2, 11.7.2}
71
Technical Summary
Future global warming exceeding the assessed very likely range cannot be ruled out and is potentially associated
with the highest risks for society and ecosystems. Such low-likelihood, high-warming storylines tend to exhibit
substantially greater changes in the intensity of regional drying and wetting than the multi-model mean. Even at
levels of warming within the very likely range, global and regional low-likelihood outcomes might occur, such as large
precipitation changes, additional sea level rise associated with collapsing ice sheets (see Box TS.4), or abrupt ocean
circulation changes. While there is medium confidence that the Atlantic Meridional Overturning Circulation (AMOC)
will not experience an abrupt collapse before 2100, if it were to occur, it would very likely cause abrupt shifts in
regional weather patterns and water cycle. The probability of these low-likelihood outcomes increases with higher
global warming levels. If the real-world climate sensitivity lies at the high end of the assessed range, then global
and regional changes substantially outside the very likely range projections occur for a given emissions scenario.
With increasing global warming, some very rare extremes and some compound events (multivariate or concurrent
extremes) with low likelihood in past and current climate will become more frequent, and there is a higher chance
that events unprecedented in the observational record occur (high confidence). Finally, low-likelihood, high-impact
outcomes may also arise from a series of very large volcanic eruptions that could substantially alter the 21st century
climate trajectory compared to SSP-based Earth system model (ESM) projections. {Cross-Chapter Box 4.1, 4.3, 4.4, 4.8,
7.3, 7.4, 7.5, 8.6, 9.2, 9.6, Box 9.4, Box 11.2, Cross-Chapter Box 12.1}
Previous IPCC reports largely focused their assessment on the projected very likely range of future surface warming and associated
climate change. However, a comprehensive risk assessment also requires considering the potentially larger changes in the physical
TS climate system that are unlikely or very unlikely but possible and potentially associated with the highest risks for society and
ecosystems (Figure TS.6). Since AR5, the development of physical climate storylines of high warming has emerged as a useful approach
for exploring the future risk space that lies outside of the IPCC very likely range projections. {4.8}
Uncertainty in the true values of equilibrium climate sensitivity (ECS) and transient climate response (TCR) dominate uncertainty
in projections of future warming under moderate to strong emissions scenarios (Section TS.3.2). A real-world ECS higher than the
assessed very likely range (2°C–5°C) would require a strong historical aerosol cooling and/or a trend towards stronger warming from
positive feedbacks linked to changes in SST patterns (pattern effects), combined with a strong positive cloud feedback and substantial
biases in paleoclimate reconstructions – each of which is assessed as either unlikely or very unlikely, but not ruled out. Since CMIP6
contains several ESMs that exceed the upper bound of the assessed very likely range in future surface warming, these models can
be used to develop low-likelihood, high warming storylines to explore risks and vulnerabilities, even in the absence of a quantitative
assessment of likelihood. {4.3.4, 4.8, 7.3.2, 7.4.4, 7.5.2, 7.5.5, 7.5.7}
CMIP6 models with surface warming outside, or close to, the upper bound of the very likely range exhibit patterns of large widespread
temperature and precipitation changes that differ substantially from the multi-model mean in all scenarios. For SSP5-8.5, the
high-warming models exhibit widespread warming of more than 6°C over most extratropical land regions and parts of the Amazon.
In the Arctic, annual mean temperatures increase by more than 10°C relative to present-day, corresponding to about 30% more than
the best estimate of warming. Even for SSP1-2.6, high-warming models show on average 2°C–3°C warming relative to present-day
conditions over much of Eurasia and North America (about 40% more than the best estimate of warming) and more than 4°C warming
relative to the present over the Arctic in 2081–2100 (Box TS.3, Figure 1). Such a high global warming storyline would imply that the
remaining carbon budget consistent with a 2°C warming is smaller than the assessed very likely range. Put another way, even if
a carbon budget that likely limits warming to 2°C is met, a low-likelihood, high-warming storyline would result in warming of 2.5°C
or more. {4.8}
CMIP6 models with global warming close to the upper bound of the assessed very likely warming range tend to exhibit greater
changes in the intensity of regional drying and wetting than the multi-model mean. Furthermore, these model projections show
a larger area of drying and tend to show a larger fraction of strong precipitation increases than the multi-model mean. However,
regional precipitation changes arise from both thermodynamic and dynamic processes so that the most pronounced global warming
levels are not necessarily associated with the strongest precipitation response. Abrupt human-caused changes to the water cycle
cannot be ruled out. Positive land surface feedbacks, involving vegetation and dust, can contribute to abrupt changes in aridity,
but there is only low confidence that such changes will occur during the 21st century. Continued Amazon deforestation, combined
with a warming climate, raises the probability that this ecosystem will cross a tipping point into a dry state during the 21st century
(low confidence). (See also Box TS.9). {4.8, 8.6.2}
72
Technical Summary
While there is medium confidence that the projected decline in the AMOC (Section TS.2.4) will not involve an abrupt collapse
before 2100, such a collapse might be triggered by an unexpected meltwater influx from the Greenland Ice Sheet. If an AMOC collapse
were to occur, it would very likely cause abrupt shifts in the regional weather patterns and water cycle, such as a southward shift in the
tropical rain belt, and could result in weakening of the African and Asian monsoons, strengthening of Southern Hemisphere monsoons,
and drying in Europe. (See also Boxes TS.9 and TS.13). {4.7.2, 8.6.1, 9.2.3}
Very rare extremes and compound or concurrent events, such as the 2018 concurrent heatwaves across the Northern Hemisphere, are
often associated with large impacts. The changing climate state is already altering the likelihood of extreme events, such as decadal
droughts and extreme sea levels, and will continue to do so under future warming. Compound events and concurrent extremes
contribute to increasing probability of low-likelihood, high-impact outcomes and will become more frequent with increasing global
warming (high confidence). Higher warming levels increase the likelihood of events unprecedented in the observational record. {9.6.4,
Box 11.2}
Finally, low likelihood storylines need not necessarily relate solely to the human-induced changes in climate. A low-likelihood, high-
impact outcome, consistent with historical precedent in the past 2500 years, would be to see several large volcanic eruptions that
could greatly alter the 21st century climate trajectory compared to SSP-based Earth system model projections. {Cross-Chapter Box 4.1}
SSP1-2.6 (2081–2100)
(a) Best estimate (scaled) (b) High-warming models (c) Very-high-warming models TS
SSP5-8.5 (2081–2100)
(d) Best estimate (scaled) (e) High-warming models (f) Very-high-warming models
–8 –6 –5 –4 –3 –2 0 2 3 4 5 6 8
°C
Box TS.3, Figure 1 | High-warming storylines. The intent of this figure is to illustrate high warming storylines compared to the CMIP6 multi-model-mean.
(a) Coupled Model Intercomparison Project Phase 6 (CMIP6) multi-model mean linearly scaled to the assessed best global surface temperature estimate for SSP1-
2.6 in 2081–2100 relative to 1995–2014, (b) mean across five high-warming models with global surface temperature changes nearest to the upper bound of the
assessed very likely range, and (c) mean across five very high-warming models with global surface temperature changes higher than the assessed very likely. (d–f)
Same as (a–c) but for SSP5-8.5. Note the different colour bars in (a–c) and (d–f). {4.7, Figure 4.41}
73
Technical Summary
TS.2.4 The Ocean Global mean SST has increased since the beginning of the 20th
century by 0.88 [0.68 to 1.01] °C, and it is virtually certain it will
Observations, models and paleo-evidence indicate that continue to increase throughout the 21st century, with increasing
recently observed changes in the ocean are unprecedented hazards to marine ecosystems (medium confidence). Marine
for centuries to millennia (high confidence). Over the past heatwaves have become more frequent over the 20th century (high
four to six decades, it is virtually certain that the global confidence), approximately doubling in frequency (high confidence)
ocean has warmed, with human influence extremely and becoming more intense and longer since the 1980s (medium
likely the main driver since the 1970s, making climate confidence). Most of the marine heatwaves over 2006–2015 have
change irreversible over centuries to millennia (medium been attributed to anthropogenic warming (very likely). Marine
confidence). It is virtually certain that upper ocean salinity heatwaves will continue to increase in frequency, with a likely global
contrasts have increased since the 1950s and extremely increase of 2–9 times in 2081–2100 compared to 1995–2014 under
likely that human influence has contributed. It is virtually SSP1-2.6, and 3–15 times under SSP5-8.5 (Figure TS.11a), with
certain that upper ocean stratification has increased since the largest changes in the tropical and Arctic ocean. {2.3.1, Cross-
1970 and that sea water pH has declined globally over the Chapter Box 2.3, 9.2.1, Box 9.2, 12.4.8}
last 40 years, with human influence being the main driver
of the observed surface open ocean acidification (virtually Observed upper-ocean stratification (0–200 m) has increased
certain). A long-term increase in surface open ocean pH globally since at least 1970 (virtually certain). Based on recent refined
occurred over the past 50 million years (high confidence), analyses of the available observations, there is high confidence
and surface ocean pH as low as recent times is uncommon that it increased by 4.9 ± 1.5% from 1970–2018, which is about
in the last 2 million years (medium confidence). There is twice as much as assessed in SROCC, and will continue to increase
high confidence that marine heatwaves have become more throughout the 21st century at a rate depending on the emissions
frequent in the 20th century, and most of those since 2006 scenario (virtually certain). {2.3.3, 9.2.1}
TS have been attributed to anthropogenic warming (very
likely). There is high confidence that oxygen levels have It is virtually certain that since 1950 near-surface high-salinity
dropped in many regions since the mid 20th century and regions have become more saline, while low-salinity regions have
that the geographic range of many marine organisms has become fresher, with medium confidence that this is linked to an
changed over the last two decades. intensification of the hydrological cycle (Box TS.6). It is extremely
likely that human influence has contributed to this salinity change
The amount of ocean warming observed since 1971 and that the large-scale pattern will grow in amplitude over the 21st
will likely at least double by 2100 under a low warming century (medium confidence). {2.3.3, 3.5.2, 9.2.2, 12.4.8}
scenario (SSP1-2.6) and will increase by 4–8 times under
a high warming scenario (SSP5-8.5). Stratification (virtually The AMOC was relatively stable during the past 8000 years (medium
certain), acidification (virtually certain), deoxygenation confidence). There is low confidence in the quantification of AMOC
(high confidence) and marine heatwave frequency (high changes in the 20th century because of low agreement in quantitative
confidence) will continue to increase in the 21st century. reconstructed and simulated trends, missing key processes in both
While there is low confidence in 20th century AMOC change, models and measurements used for formulating proxies, and new
it is very likely that AMOC will decline over the 21st century model evaluations. Direct observational records since the mid-2000s
(Figure TS.11). {2.3, 3.5, 3.6, 4.3.2, 5.3, 7.2, 9.2, Box 9.2, 12.4} are too short to determine the relative contributions of internal
variability, natural forcing and anthropogenic forcing to AMOC
It is virtually certain that the global ocean has warmed since at least change (high confidence). An AMOC decline over the 21st century
1971, representing about 90% of the increase in the global energy is very likely for all SSP scenarios (Figure TS.11b); a possible abrupt
inventory (Section TS.3.1). The ocean is currently warming faster than decline is assessed further in Box TS.3. {2.3.3, 3.5.4, 4.3.2, 8.6.1,
at any other time since at least the last deglacial transition (medium 9.2.3, Cross-Chapter Box 12.3}
confidence), with warming extending to depths well below 2000 m
(very high confidence). It is extremely likely that human influence There is high confidence that many ocean currents will change in
was the main driver of this recent ocean warming. Ocean warming the 21st century in response to changes in wind stress. There is low
will continue over the 21st century (virtually certain), and will likely confidence in 21st century change of Southern Ocean circulation,
continue until at least to 2300 even for low CO2 emissions scenarios. despite high confidence that it is sensitive to changes in wind
Ocean warming is irreversible over centuries to millennia (medium patterns and increased ice-shelf melt. Western boundary currents
confidence), but the magnitude of warming is scenario-dependent from and subtropical gyres have shifted poleward since 1993 (medium
about the mid-21st century (medium confidence). The warming will confidence). Subtropical gyres, the East Australian Current Extension,
not be globally uniform, with heat primarily stored in Southern Ocean the Agulhas Current, and the Brazil Current are projected to intensify
water-masses and weaker warming in the subpolar North Atlantic (high in the 21st century in response to changes in wind stress, while the
confidence). Limitations in the understanding of feedback mechanisms Gulf Stream and the Indonesian Throughflow are projected to weaken
limit our confidence in future ocean warming close to Antarctica and (medium confidence). All of the four main eastern boundary upwelling
how this will affect sea ice and ice shelves. {2.3.3, 3.5.1, 4.7.2, 7.2.2, systems are projected to weaken at low latitudes and intensify at high
9.2.2, 9.2.3, 9.2.4, 9.3.2, 9.6.1, Cross-Chapter Box 9.1} latitudes in the 21st century (high confidence). {2.3.3, 9.2.3}
74
Technical Summary
10
% change
-5
5 Past (observed) Future (CMIP6)
-10
Past (simulated)
-15
Change Change
in 2100 in 2100
pH
Sv
-5
7.8
high acidity Future (CMIP6)
-10 Future (CMIP6)
7.6
Change Change
1850 1900 1950 2000 2050 2100 in 2100 1850 1900 1950 2000 2050 2100 in 2100
(1978–2017) (2061–2100)
m
m
104 Gt
-4
1 0.1 -8 1 0.2
Elevation change
Elevation change
0.5 0.5
0.3
(m/yr)
(m/yr)
0 0
-0.5 0.2 -12
-0.5
-8 -1 -1
0.4
ISMIP6 Emulator -16
1980 2000 2020 2040 2060 2080 2100 Change 1980 2000 2020 2040 2060 2080 2100
in 2100
ISMIP6 LARMIP
Observation-based: Emulator
Emulator median (SSP1-2.6); 17-83% & 5–95% ranges
Bamber Change
Emulator median (SSP5-8.5); 17-83% & 5–95% ranges in 2100
Mouginot/Rignot ISMIP6 models (SSP1-2.6/RCP2.6)
IMBIE ISMIP6 models (SSP5-8.5/RCP8.5)
Figure TS.11 | Past and future ocean and ice-sheet changes. The intent of this figure is to show that observed and projected time series of many ocean and cryosphere
indicators are consistent. Observed and simulated historical changes and projected future changes under varying greenhouse gas emissions scenarios. Simulated and projected
ocean changes are shown as Coupled Model Intercomparison Project Phase 6 (CMIP6) ensemble mean, and 5–95% range (shading) is provided for scenarios SSP1-2.6 and
SSP3-7.0 (except in panel a where the range is provided for scenario SSP1-2.6 and SSP5-8.5). Mean and 5–95% range in 2100 are shown as vertical bars on the right-hand side
of each panel. (a) Change in multiplication factor in surface ocean marine heatwave days relative to 1995–2014 (defined as days exceeding the 99th percentile in sea surface
temperature (SST) from 1995–2014 distribution). Assessed observational change span 1982–2019 from AVHRR satellite SST. (b) Atlantic Meridional Overturning Circulation
(AMOC) transport relative to 1995–2014 (defined as maximum transport at 26°N). Assessed observational change spans 2004–2018 from the RAPID array smoothed with
a 12-month running mean (shading around the mean shows the 12-month running standard deviation around the mean). (c) Global mean percent change in ocean oxygen
(100–600 m depth), relative to 1995–2014. Assessed observational trends and very likely range are from the SROCC assessment, and span 1970–2010 centred on 2005. (d)
Global mean surface pH. Assessed observational change spans 1985–2019, from the CMEMS SOCAT-based reconstruction (shading around the global mean shows the 90%
confidence interval). (e), (f): Ice sheet mass changes. Projected ice-sheet changes are shown as median, 5–95% range (light shading), and 17–83% range (dark shading) of
cumulative mass loss and sea level equivalent from ISMIP6 emulation under SSP1-2.6 and SSP5-8.5 (shading and bold line), with individual emulated projections as thin lines.
Median (dot), 17–83% range (thick vertical bar), and 5–95% range (thin vertical bar) in 2100 are shown as vertical bars on the right-hand side of each panel, from ISMIP6,
ISMIP6 emulation, and LARMIP-2. Observation-based estimates: For Greenland (e), for 1972–2018 (Mouginot), for 1992–2016 (Bamber), for 1992–2020 (IMBIE) and total
estimated mass loss range for 1840–1972 (Box). For Antarctica (f), estimates based on satellite data combined with simulated surface mass balance and glacial isostatic
adjustment for 1992–2020 (IMBIE), 1992–2016 (Bamber), and 1979–2017 (Rignot). Left inset maps: mean Greenland elevation changes 2010–2017 derived from CryoSat-2
radar altimetry (e) and mean Antarctica elevation changes 1978–2017 derived from restored analogue radar records (f). Right inset maps: ISMIP6 model mean (2093–2100)
projected changes under the MIROC5 climate model for the RCP8.5 scenario. {2.3.3; 2.3.4; 3.5.4; 4.3.2; 5.3.2; 5.3.3; 5.6.3; 9.2.3; 9.4.1; 9.4.2; Box 9.2; Box 9.2, Figure 1;
Figures 9.10, 9.17 and 9.18}
75
Technical Summary
It is virtually certain that surface pH has declined globally over the become practically sea ice-free in late summer under high
last 40 years and that the main driver is uptake of anthropogenic CO2 emissions scenarios by the end of the 21st century (high
CO2. Ocean acidification and associated reductions in the saturation confidence). It is virtually certain that further warming will
state of calcium carbonate – a constituent of skeletons or shells of lead to further reductions of Northern Hemisphere snow
a variety of marine organisms – is expected to increase in the 21st cover, and there is high confidence that this is also the case
century under all emissions scenarios (high confidence). A long-term for near-surface permafrost volume.
increase in surface open ocean pH occurred over the past 50 million
years (high confidence), and surface ocean pH as low as recent Glaciers will continue to lose mass at least for several
times is uncommon in the last 2 million years (medium confidence). decades even if global temperature is stabilized (very
There is very high confidence that present-day surface pH values high confidence), and mass loss over the 21st century
are unprecedented for at least 26,000 years and current rates of pH is virtually certain for the Greenland Ice Sheet and likely
change are unprecedented since at least that time. Over the past 2–3 for the Antarctic Ice Sheet. Deep uncertainty persists with
decades, a pH decline in the ocean interior has been observed in all respect to the possible evolution of the Antarctic Ice Sheet
ocean basins (high confidence) (Figure TS.11d). {2.3.3, 2.3.4, 3.6.2, within the 21st century and beyond, in particular due to the
4.3.2, 5.3.2, 5.3.3, 5.6.3, 12.4.8} potential instability of the West Antarctic Ice Sheet. {2.3,
3.4, 4.3, 8.3, 9.3–9.6, Box 9.4, 12.4}
Open-ocean deoxygenation and expansion of oxygen minimum zones
have been observed in many areas of the global ocean since the Current Arctic sea ice coverage levels (both annual and late summer)
mid 20th century (high confidence), in part due to human influence are at their lowest since at least 1850 (high confidence), and for
(medium confidence). Deoxygenation is projected to continue to late summer for the past 1000 years (medium confidence). Since the
increase with ocean warming (high confidence) (Figure TS.11c). late 1970s, Arctic sea ice area and thickness have decreased in both
Higher climate sensitivity and reduced ocean ventilation in CMIP6 summer and winter, with sea ice becoming younger, thinner and more
TS compared to CMIP5 results in substantially greater projections of dynamic (very high confidence). It is very likely that anthropogenic
subsurface (100–600 m) oxygen decline than reported in SROCC for forcing, mainly due to greenhouse gas increases, was the main driver
the period 2080–2099. {2.3.3, 2.3.4, Cross-Chapter Box 2.4, 3.6.2, of this loss, although new evidence suggests that anthropogenic
5.3.3, 12.4.8} aerosol forcing has offset part of the greenhouse gas-induced losses
since the 1950s (medium confidence). The annual Arctic sea ice area
Over at least the last two decades, the geographic range of many minimum will likely fall below 1 million km2 at least once before
marine organisms has shifted towards the poles and towards greater 2050 under all assessed SSP scenarios. This practically sea ice-free
depths (high confidence), indicative of shifts towards cooler waters. state will become the norm for late summer by the end of the 21st
The range of a smaller subset of organisms has shifted equatorward century in high CO2 emissions scenarios (high confidence). Arctic
and to shallower depths (high confidence). Phenological metrics summer sea ice varies approximately linearly with global surface
associated with the life cycles of many organisms have also temperature, implying that there is no tipping point and observed/
changed over the last two decades or longer (high confidence). projected losses are potentially reversible (high confidence). {2.3.2,
Since the changes in the geographical range of organisms and their 3.4.1, 4.3.2, 9.3.1, 12.4.9}
phenological metrics have been observed to differ with species
and location, there is the possibility of disruption to major marine For Antarctic sea ice, there is no significant trend in satellite-observed
ecosystems. {2.3.4} sea ice area from 1979 to 2020 in both winter and summer, due to
regionally opposing trends and large internal variability. Due to
mismatches between model simulations and observations, combined
TS.2.5 The Cryosphere with a lack of understanding of reasons for substantial inter-model
spread, there is low confidence in model projections of future Antarctic
Over recent decades, widespread loss of snow and ice has sea ice changes, particularly at the regional level. {2.3.2, 3.4.1, 9.3.2}
been observed, and several elements of the cryosphere
are now in states unseen in centuries (high confidence). In permafrost regions, increases in ground temperatures in the upper
Human influence was very likely the main driver of 30 m over the past three to four decades have been widespread (high
observed reductions in Arctic sea ice since the late 1970s confidence). For each additional 1°C of warming (up to 4°C above the
(with late-summer sea ice loss likely unprecedented for at 1850–1900 level), the global volume of perennially frozen ground to
least 1000 years) and the widespread retreat of glaciers 3 m below the surface is projected to decrease by about 25% relative
(unprecedented in at least the last 2,000 years, medium to the present volume (medium confidence). However, these decreases
confidence). Furthermore, human influence very likely may be underestimated due to an incomplete representation of
contributed to the observed Northern Hemisphere spring relevant physical processes in ESMs (low confidence). Seasonal snow
snow cover decrease since 1950. cover is treated in Section TS.2.6. {2.3.2, 9.5.2, 12.4.9}
By contrast, Antarctic sea ice area experienced no significant There is very high confidence that, with few exceptions, glaciers
net change since 1979, and there is only low confidence have retreated since the second half of the 19th century; this
in its projected changes. The Arctic Ocean is projected to behaviour is unprecedented in at least the last 2000 years (medium
76
Technical Summary
confidence). Mountain glaciers very likely contributed 67.2 [41.8 to of the Antarctic Ice Sheet, is assessed in Box TS.9. {2.3.2, 3.4.3, 9.4.1,
92.6] mm to the observed GMSL change between 1901 and 2018. 9.4.2, 9.6.3, Atlas.11.2}
This retreat has occurred at increased rates since the 1990s, with
human influence very likely being the main driver. Under RCP2.6 and It is likely that the Antarctic Ice Sheet has lost 2670 ± 530 Gt,
RCP8.5, respectively, glaciers are projected to lose 18% ± 13% and contributing 7.4 ± 1.5 mm to GMSL rise over 1992–2020. The total
36% ± 20% of their current mass over the 21st century (medium Antarctic ice mass losses were dominated by the West Antarctic Ice
confidence). {2.3.2, 3.4.3, 9.5.1, 9.6.1} Sheet, with combined West Antarctic and Peninsula annual loss rates
increasing since about 2000 (very high confidence). Furthermore, it
The Greenland Ice Sheet was smaller than at present during the is very likely that parts of the East Antarctic Ice Sheet have lost mass
Last Interglacial period (roughly 125,000 years ago) and the mid- since 1979. Since the 1970s, snowfall has likely increased over the
Holocene (roughly 6,000 years ago) (high confidence). After reaching western Antarctic Peninsula and eastern West Antarctica, with large
a recent maximum ice mass at some point between 1450 and 1850, spatial and interannual variability over the rest of Antarctica. Mass
the ice sheet retreated overall, with some decades likely close to losses from West Antarctic outlet glaciers, mainly induced by ice shelf
equilibrium (i.e., mass loss approximately equalling mass gained). It basal melt (high confidence), outpace mass gain from increased
is virtually certain that the Greenland Ice Sheet has lost mass since snow accumulation on the continent (very high confidence).
the 1990s, with human influence a contributing factor (medium However, there is only limited evidence, with medium agreement,
confidence). There is high confidence that annual mass changes have of anthropogenic forcing of the observed Antarctic mass loss since
been consistently negative since the early 2000s. Over the period 1992 (with low confidence in process attribution). Increasing mass
1992–2020, Greenland likely lost 4890 ± 460 Gt of ice, contributing loss from ice shelves and inland discharge will likely continue to
13.5 ± 1.3 mm to GMSL rise. There is high confidence that Greenland outpace increasing snowfall over the 21st century (Figure TS.11f).
ice mass losses are increasingly dominated by surface melting and Deep uncertainty persists with respect to the possible evolution of
runoff, with large interannual variability arising from changes in the Antarctic Ice Sheet along high-end mass-loss storylines within
surface mass balance. Projections of future Greenland ice-mass loss the 21st century and beyond, primarily related to the abrupt and TS
(Box TS.4, Table 1; Figure TS.11e) are dominated by increased surface widespread onset of marine ice sheet instability and marine ice cliff
melt under all emissions scenarios (high confidence). Potential instability. (See also Boxes TS.3 and TS.4). {2.3.2, 3.4.3, 9.4.2, 9.6.3,
irreversible long-term loss of the Greenland Ice Sheet, and of parts Box 9.4, Atlas.11.1}
Global mean sea level (GMSL) increased by 0.20 [0.15 to 0.25] m over the period 1901 to 2018, with a rate of rise
that has accelerated since the 1960s to 3.7 [3.2 to 4.2] mm yr –1 for the period 2006–2018 (high confidence). Human
activities were very likely the main driver of observed GMSL rise since 1971, and new observational evidence leads
to an assessed sea level rise over the period 1901 to 2018 that is consistent with the sum of individual components
contributing to sea level rise, including expansion due to ocean warming and melting of glaciers and ice sheets (high
confidence). It is virtually certain that GMSL will continue to rise over the 21st century in response to continued
warming of the climate system (Box TS.4, Figure 1). Sea level responds to greenhouse gas (GHG) emissions more
slowly than global surface temperature, leading to weaker scenario dependence over the 21st century than for global
surface temperature (high confidence). This slow response also leads to long-term committed sea level rise, associated
with ongoing ocean heat uptake and the slow adjustment of the ice sheets, that will continue over the centuries and
millennia following cessation of emissions (high confidence) (Box TS.9). By 2100, GMSL is projected to rise by 0.28–
0.55 m (likely range) under SSP1-1.9 and 0.63–1.01 m (likely range) under SSP5-8.5 relative to the 1995–2014 average
(medium confidence). Under the higher CO2 emissions scenarios, there is deep uncertainty in sea level projections for
2100 and beyond associated with the ice-sheet responses to warming. In a low-likelihood, high-impact storyline and a
high CO2 emissions scenario, ice-sheet processes characterized by deep uncertainty could drive GMSL rise up to about
5 m by 2150. Given the long-term commitment, uncertainty in the timing of reaching different GMSL rise levels is an
important consideration for adaptation planning. {2.3, 3.4, 3.5, 9.6, Box 9.4, Cross-Chapter Box 9.1, Table 9.5}
GMSL change is driven by warming or cooling of the ocean (and the associated expansion/contraction) and changes in the amount
of ice and water stored on land. Paleo-evidence shows that GMSL has been about 70 m higher and 130 m lower than present within
the past 55 million years and was likely 5 to 10 m higher during the Last Interglacial (Box TS.2, Figure 1). Sea level observations show
that GMSL rose by 0.20 [0.15 to 0.25] m over the period 1901–2018 at an average rate of 1.7 [1.3 to 2.2] mm yr –1. New analyses
and paleo-evidence since AR5 show this rate is very likely faster than during any century over at least the last three millennia (high
confidence). Since AR5, there is strengthened evidence for an increase in the rate of GMSL rise since the mid-20th century, with an
average rate of 2.3 [1.6–3.1] mm yr –1 over the period 1971–2018 increasing to 3.7 [3.2–4.2] mm yr –1 for the period 2006–2018 (high
confidence). {2.3.3, 9.6.1, 9.6.2}
77
Technical Summary
0.5
Observations SSP1-2.6 2150 medium & low
0 confidence projections
(see caption)
1900 1950 2000 2050 2100 2150
(b) Committed sea level rise by warming level and time scale (c) Projected timing of sea level rise milestones
Paleo ranges
35
TS Year by which a rise of 2.0 m
above 1995–2014 is expected
30 10,000-yr
Mid-Pliocene
Warm Period
25
GMSL rise (m)
1.5 m
20
15 2,000-yr
Last
Interglacial 1.0 m
10
Low−confidence
processes included SSP5−8.5
SSP3−7.0
5 SSP2−4.5
SSP1−2.6
SSP1−1.9
100-yr 0.5 m
0
1 1.5 2 3 4 5
2000 2100 2200 2300+
Peak global surface temperature (°C)
Box TS.4, Figure 1 | Global mean sea level (GMSL) change on different time scales and under different scenarios. The intent of this figure is to
(i) show the century-scale GMSL projections in the context of the 20th century observations, (ii) illustrate ‘deep uncertainty’ in projections by considering the timing
of GMSL rise milestones, and (iii) show the long-term commitment associated with different warming levels, including the paleo evidence to support this. (a) GMSL
change from 1900 to 2150, observed (1900–2018) and projected under the SSP scenarios (2000–2150), relative to a 1995–2014 baseline. Solid lines show median
projections. Shaded regions show likely ranges for SSP1-2.6 and SSP3-7.0. Dotted and dashed lines show respectively the 83rd and 95th percentile low confidence
projections for SSP5-8.5. Bars at right show likely ranges for SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0 and SSP5-8.5 in 2150. Lightly shaded thick/thin bars show
17th–83rd/5th–95th percentile low-confidence ranges in 2150 for SSP1-2.6 and SSP5-8.5, based upon projection methods incorporating structured expert judgement
and marine ice cliff instability. Low confidence range for SSP5-8.5 in 2150 extends to 4.8/5.4 m at the 83rd/95th percentile. (b) GMSL change on 100- (blue), 2000-
(green) and 10,000-year (magenta) time scales as a function of global surface temperature, relative to 1850–1900. For 100-year projections, GMSL is projected for the
year 2100, relative to a 1995–2014 baseline, and temperature anomalies are average values over 2081–2100. For longer-term commitments, warming is indexed by
peak warming above 1850–1900 reached after cessation of emissions. Shaded regions show paleo-constraints on global surface temperature and GMSL for the Last
Interglacial and mid-Pliocene Warm Period. Lightly shaded thick/thin blue bars show 17th–83rd/5th–95th percentile low confidence ranges for SSP1-2.6 and SSP5-8.5
in 2100, plotted at 2°C and 5°C. (c) Timing of exceedance of GMSL thresholds of 0.5, 1.0, 1.5 and 2.0 m, under different SSPs. Lightly shaded thick/thin bars show
17th–83rd/5th–95th percentile low-confidence ranges for SSP1-2.6 and SSP5-8.5. {4.3.2, 9.6.1, 9.6.2, 9.6.3, Box 9.4}
78
Technical Summary
GMSL will continue to rise throughout the 21st century (Box TS.4, Figure 1a). Considering only those processes in whose projections
we have at least medium confidence, relative to the period 1995–2014, GMSL is projected to rise between 0.18 m (0.15–0.23 m,
likely range; SSP1-1.9) and 0.23 m (0.20–0.30 m, likely range; SSP5-8.5) by 2050. By 2100, the projected rise is between 0.38 m
(0.28–0.55 m, likely range; SSP1-1.9) and 0.77 m (0.63–1.01 m, likely range; SSP5-8.5) {Table 9.9}. The methods, models and scenarios
used for sea level projections in the AR6 are updated from those employed by SROCC, with contributions informed by the latest
model projections described in the ocean and cryosphere Sections (Sections TS.2.4 and TS.2.5). Despite these differences, the sea level
projections are broadly consistent with those of SROCC. {4.3.2, 9.6.3}
Importantly, likely range projections do not include those ice-sheet-related processes whose quantification is highly uncertain
or that are characterized by deep uncertainty. Higher amounts of GMSL rise before 2100 could be caused by earlier-than-projected
disintegration of marine ice shelves, the abrupt, widespread onset of marine ice sheet instability (MISI) and marine ice cliff instability
(MICI) around Antarctica, and faster-than-projected changes in the surface mass balance and dynamical ice loss from Greenland
(Box TS.4, Figure 1). In a low-likelihood, high-impact storyline and a high CO2 emissions scenario, such processes could in combination
contribute more than one additional meter of sea level rise by 2100 (Box TS.3). {4.3.2, 9.6.3, Box 9.4}
Beyond 2100, GMSL will continue to rise for centuries to millennia due to continuing deep ocean heat uptake and mass loss from
ice sheets, and will remain elevated for thousands of years (high confidence). By 2150, considering only those processes in whose
projections we have at least medium confidence and assuming no acceleration in ice-mass flux after 2100, GMSL is projected to rise
between 0.6 m (0.4–0.9 m, likely range, SSP1-1.9) and 1.3 m (1.0–1.9 m, likely range) (SSP5-8.5), relative to the period 1995–2014
based on the SSP scenario extensions. Under high CO2 emissions, processes in which there is low confidence, such as MICI, could drive TS
GMSL rise up to about 5 m by 2150 (Box TS.4, Figure 1a). By 2300, GMSL will rise 0.3–3.1 m under low CO2 emissions (SSP1-2.6) (low
confidence). Under high CO2 emissions (SSP5-8.5), projected GMSL rise is between 1.7 and 6.8 m by 2300 in the absence of MICI and
by up to 16 m considering MICI (low confidence). Over 2000 years, there is medium agreement and limited evidence that committed
GMSL rise is projected to be about 2–3 m with 1.5°C peak warming, 2–6 m with 2°C of peak warming, 4–10 m with 3°C of peak
warming, 12–16 m with 4°C of peak warming, and 19–22 m with 5°C of peak warming. {9.6.3}
Looking at uncertainty in time provides an alternative perspective on uncertainty in future sea level rise (Box TS.4, Figure 1c). For
example, considering only medium confidence processes, GMSL rise is likely to exceed 0.5 m between about 2080 and 2170 under
SSP1-2.6 and between about 2070 and 2090 under SSP5-8.5. Given the long-term commitment, uncertainty in the timing of reaching
different levels of GMSL rise is an important consideration for adaptation planning. {9.6.3}
At regional scales, additional processes come into play that modify the local sea level change relative to GMSL, including vertical land
motion, ocean circulation and density changes, and gravitational, rotational, and deformational effects arising from the redistribution
of water and ice mass between land and the ocean. These processes give rise to a spatial pattern that tends to increase sea level
rise at the low latitudes and reduce sea level rise at high latitudes. However, over the 21st century, the majority of coastal locations
have a median projected regional sea level rise within ±20% of the projected GMSL change (medium confidence). Further details on
regional sea level change and extremes are provided in Section TS.4. {9.6.3}
The continued growth of atmospheric CO2 concentrations over the industrial era is unequivocally due to emissions from
human activities. Ocean and land carbon sinks slow the rise of CO2 in the atmosphere. Projections show that while
land and ocean sinks absorb more CO2 under high emissions scenarios than low emissions scenarios, the fraction of
emissions removed from the atmosphere by natural sinks decreases with higher concentrations (high confidence).
Projected ocean and land sinks show similar responses for a given scenario, but the land sink has a much higher
interannual variability and wider model spread. The slowed growth rates of the carbon sinks projected for the second
half of this century are linked to strengthening carbon–climate feedbacks and stabilization of atmospheric CO2 under
medium-to-no-mitigation and high-mitigation scenarios, respectively (see FAQ 5.1). {5.2, 5.4}
79
Technical Summary
Carbon sinks for anthropogenic CO2 are associated with mainly physical ocean and biospheric land processes that drive the exchange
of carbon between multiple land, ocean and atmospheric reservoirs. These exchanges are driven by increasing atmospheric CO2, but
are modulated by changes in climate (Box TS.5, Figure 1c,d). The Northern and Southern Hemispheres dominate the land and ocean
sinks, respectively (Box TS.5, Figure 1). Ocean circulation and thermodynamic processes also play a critical role in coupling the global
carbon and energy (heat) cycles. There is high confidence that this ocean carbon–heat nexus is an important basis for one of the
most important carbon–climate metrics, the transient climate response to cumulative CO2 emissions (TCRE; Section TS.3.2.1) used to
determine the remaining carbon budget. {5.1, 5.2, 5.5, 9.2, Cross-Chapter Box 5.3}
Based on multiple lines of evidence using interhemispheric gradients of CO2 concentrations, isotopes, and inventory data, it is
unequivocal that the growth in CO2 in the atmosphere since 1750 (see Section TS.2.2) is due to the direct emissions from human
activities. The combustion of fossil fuels and land-use change for the period 1750–2019 resulted in the release of 700 ± 75 PgC (likely
range, 1 PgC = 1015 g of carbon) to the atmosphere, of which about 41% ± 11% remains in the atmosphere today (high confidence).
Of the total anthropogenic CO2 emissions, the combustion of fossil fuels was responsible for about 64% ± 15%, growing to an 86% ±
14% contribution over the past 10 years. The remainder resulted from land-use change. During the last decade (2010–2019), average
annual anthropogenic CO2 emissions reached the highest levels in human history at 10.9 ± 0.9 PgC yr –1 (high confidence). Of these
emissions, 46% accumulated in the atmosphere (5.1 ± 0.02 PgC yr –1), 23% (2.5 ± 0.6 PgC yr –1) was taken up by the ocean and 31%
(3.4 ± 0.9 PgC yr –1) was removed by terrestrial ecosystems (high confidence). {5.2.1, 5.2.2, 5.2.3}
The ocean (high confidence) and land (medium confidence) sinks of CO2 have increased with anthropogenic emissions over the past
TS six decades (Box TS.5, Figure 1). This coherence between emissions and the growth in ocean and land sinks has resulted in the airborne
fraction of anthropogenic CO2 remaining at 44 ± 10% over the past 60 years (high confidence). Interannual and decadal variability of
the ocean and land sinks indicate that they are sensitive to changes in the growth rate of emissions as well as climate variability and
are therefore also sensitive to climate change (high confidence). {5.2.1}
The land CO2 sink is driven by carbon uptake by vegetation, with large interannual variability, for example, linked to the El Niño–
Southern Oscillation (ENSO). Since the 1980s, carbon fertilization from rising atmospheric CO2 has increased the strength of the net
land CO2 sink (medium confidence). During the historical period, the growth of the ocean sink has been primarily determined by
the growth rate of atmospheric CO2. However, there is medium confidence that changes to physical and chemical processes in the
ocean and in the land biosphere, which govern carbon feedbacks, are already modifying the characteristics of variability, particularly
the seasonal cycle of CO2, in both the ocean and land. However, changes to the multi-decadal trends in the sinks have not yet been
observed. {2.3.4, 3.6.1, 5.2.1}
In AR6, ESM projections are assessed with CO2 concentrations by 2100 from about 400 ppm (SSP1-1.9) to above 1100 ppm (SSP5-8.5).
Most simulations are performed with prescribed atmospheric CO2 concentrations, which already account for a central estimate of
climate–carbon feedback effects. Carbon dioxide emissions-driven simulations account for uncertainty in these feedbacks, but do
not significantly change the projected global surface temperature changes (high confidence). Although land and ocean sinks absorb
more CO2 under high emissions than low emissions scenarios, the fraction of emissions removed from the atmosphere decreases
(high confidence). This means that the more CO2 that is emitted, the less efficient the ocean and land sinks become (high confidence),
an effect which compensates for the logarithmic relationship between CO2 and its radiative forcing, which means that for each unit
increase in additional atmospheric CO2 the effect on global temperature decreases. (Box TS.5, Figure 1f,g). {4.3.1, 5.4.5, 5.5.1.2}
Ocean and land sinks show similar responses for a given scenario, but the land sink has a much higher interannual variability
and wider model spread. Under SSP3-7.0 and SSP5-8.5, the initial growth of both sinks in response to increasing atmospheric
concentrations of CO2 is subsequently limited by emerging carbon–climate feedbacks (high confidence) (Box TS.5, Figure 1f).
Projections show that the ocean and land sinks will stop growing from the second part of the 21st century under all emissions
scenarios, but with different drivers for different emissions scenarios. Under SSP3-7.0 and SSP5-8.5, the weakening growth rate
of the ocean CO2 sink in the second half of the century is primarily linked to the strengthening positive feedback from reduced
carbonate buffering capacity, ocean warming and altered ocean circulation (e.g., AMOC changes). In contrast, for SSP1-1.9, SSP1-2.6
and SSP2-4.5, the weakening growth rate of the ocean carbon sink is a response to the stabilizing or declining atmospheric
CO2 concentrations. Under SSP1-1.9, models project that combined land and ocean sinks will turn into a weak source by 2100
(medium confidence). Under high CO2 emissions scenarios, it is very likely that the land carbon sink will grow more slowly due to
warming and drying from the mid-21st century, but it is very unlikely that it will switch from being a sink to a source before 2100.
80
Technical Summary
Latitute
Colour High model agreement Low model agreement
TS
81
Technical Summary
Climate change alone is expected to increase land carbon accumulation in the high latitudes (not including permafrost, which is
assessed in Sections TS.2.5 and TS.3.2.2), but also to lead to a counteracting loss of land carbon in the tropics (medium confidence).
Earth system model projections show that the overall uncertainty of atmospheric CO2 by 2100 is still dominated by the emissions
pathway, but carbon–climate feedbacks (see Section TS.3.3.2) are important, with increasing uncertainties in high emissions pathways
(Box TS.5, Figure 1e). {4.3.2, 5.4.1, 5.4.2, 5.4.4, 5.4.5, 11.6, 11.9, Cross-Chapter Box 5.1, Cross-Chapter Box 5.3}
Under three SSP scenarios with long-term extensions until 2300 (SSP5-8.5, SSP5-3.4-OS, SSP1-2.6), ESMs project a change of the
land from a sink to a source (medium confidence). The scenarios make simplified assumptions about emissions reductions, with
SSP1-2.6 and SSP5-3.4-OS reaching about 400 ppm by 2300, while SSP5-8.5 exceeds 2000 ppm. Under high emissions, the transition
is warming-driven, whereas it is linked to the decline in atmospheric CO2 under net negative CO2 emissions. The ocean remains a sink
throughout the period to 2300 except under very large net negative emissions. The response of the natural aspects of the carbon cycle
TS to carbon dioxide removal is further developed in Section TS.3.3.2. {5.4.9}
TS.2.6 Land Climate, Including Biosphere and Extremes and about 80% larger than warming of the ocean surface. Warming of
the land surface during the period 1971–2018 contributed about 5%
Land surface air temperatures have risen faster than the of the increase in the global energy inventory (Section TS.3.1), nearly
global surface temperature since the 1850s, and it is virtually twice the estimate in AR5 (high confidence). It is virtually certain that
certain that this differential warming will persist into the the average surface warming over land will continue to be higher than
future. It is virtually certain that the frequency and intensity over the ocean throughout the 21st century. The warming pattern
of hot extremes and the intensity and duration of heatwaves will likely vary seasonally, with northern high latitudes warming more
have increased since 1950 and will further increase in the during winter than summer (medium confidence). {2.3.1, 4.3.1, 4.5.1,
future even if global warming is stabilized at 1.5°C. The 7.2.2, Box 7.2, Cross-Chapter Box 9.1, 11.3, Atlas 11.2}
frequency and intensity of heavy precipitation events have
increased over a majority of those land regions with good The frequency and intensity of hot extremes (warm days and nights)
observational coverage (high confidence) and will extremely and the intensity and duration of heatwaves have increased globally
likely increase over most land regions with additional global and in most regions since 1950, while the frequency and intensity
warming. of cold extremes have decreased (virtually certain). There is high
confidence that the increases in frequency and severity of hot
Over the past half century, key aspects of the biosphere extremes are due to human-induced climate change. Some recent
have changed in ways that are consistent with large- extreme events would have been extremely unlikely to occur without
scale warming: climate zones have shifted poleward, and human influence on the climate system. It is virtually certain that
the growing season length in the Northern Hemisphere further changes in hot and cold extremes will occur throughout the
extratropics has increased (high confidence). The amplitude 21st century in nearly all inhabited regions, even if global warming
of the seasonal cycle of atmospheric CO2 poleward of 45°N is stabilized at 1.5°C (Table TS.2, Figure TS.12a). {1.3, Cross-Chapter
has increased since the 1960s (very high confidence), with Box 3.2, 11.1.4, 11.3.2, 11.3.4, 11.3.5, 11.9, 12.4}
increasing productivity of the land biosphere due to the
increasing atmospheric CO2 concentration as the main driver Greater warming over land alters key water cycle characteristics
(medium confidence). Global-scale vegetation greenness has (Box TS.6). The rates of change in mean precipitation and runoff,
increased since the 1980s (high confidence). {2.3, 3.6, 4.3, and their variability, increase with global warming (Figure TS.12e,f).
4.5, 5.2, 11.3, 11.4, 11.9, 12.4} Human-induced climate change has contributed to increases in
agricultural and ecological droughts in some regions due to increases
Observed temperatures over land have increased by 1.59 [1.34– in evapotranspiration (medium confidence). More regions are
1.83] °C between the period 1850–1900 and 2011–2020. Warming affected by increases in agricultural and ecological droughts with
of the land is about 45% larger than for global surface temperature increasing global warming (high confidence; see also Figure TS.12c).
82
Technical Summary
(c) Droughts in drought-prone regions (d) Northern Hemisphere March–May snow cover
8
(e) Hydrological change over tropical land (f) Hydrological change over extratropical land
50 Precipitable water annual mean Precipitable water annual mean
Precipitation 60 Precipitation
% change - 15 models mean - SSP5-8.5
20 30
10 20
10
0
17–83% range 0 17–83% range
-10
2.0 3.0 4.0 5.0 2.0 3.0 4.0 5.0
Global surface temperature change since 1850–1900 (°C) Global surface temperature change since 1850–1900 (°C)
Figure TS.12 | Land-related changes relative to the 1850-1900 as a function of global warming levels. The intent of this figure is to show that extremes and
mean land variables change consistently with warming levels and to show the changes with global warming levels of water cycle indicators (i.e., precipitation and runoff) over
tropical and extratropical land in terms of mean and interannual variability (interannual variability increases at a faster rate than the mean). (a) Changes in the frequency (left
scale) and intensity (in °C, right scale) of daily hot extremes occurring every 10 and 50 years. (b) as (a), but for daily heavy precipitation extremes, with intensity change in %.
(c) Changes in 10-year droughts aggregated over drought-prone regions (WNA, CNA, NCA, SCA, NSA, NES, SAM, SWS, SSA, WCE, MED, WSAF, ESAF, MDG, SAU, and EAU;
for definitions of these regions, see Figure Atlas.2), with drought intensity (right scale) represented by the change of annual mean soil moisture, normalized with respect to
interannual variability. Limits of the 5%−95% confidence interval are shown in panels (a–c). (d) Changes in Northern Hemisphere spring (March–April–May) snow cover extent
relative to 1850–1900; (e,f) Relative change (%) in annual mean of total precipitable water (grey line), precipitation (red solid lines), runoff (blue solid lines) and in standard
deviation (i.e., variability) of precipitation (red dashed lines) and runoff (blue dashed lines) averaged over (e) tropical and (f) extratropical land as function of global warming
levels. Coupled Model Intercomparison Project Phase 6 (CMIP6) models that reached a 5°C warming level above the 1850–1900 average in the 21st century in SSP5-8.5 have
been used. Precipitation and runoff variability are estimated by respective standard deviation after removing linear trends. Error bars show the 17–83% confidence interval for
the warmest +5°C global warming level. {Figures 8.16, 9.24, 11.6, 11.7, 11.12, 11.15, 11.18 and Atlas.2}
83
Technical Summary
There is low confidence that the increase of plant water-use efficiency Hemisphere extratropics since the mid-20th century (high
due to higher atmospheric CO2 concentration alleviates extreme confidence), are consistent with large-scale warming. At the
agricultural and ecological droughts in conditions characterized same time an increase in the amplitude of the seasonal cycle of
by limited soil moisture and increased atmospheric evaporative atmospheric CO2 poleward of 45°N since the early 1960s (high
demand. {2.3.1, Cross-Chapter Box 5.1, 8.2.3, 8.4.1, 11.2.4, 11.4, confidence) and a global-scale increase in vegetation greenness
11.6, Box 11.1} of the terrestrial surface since the early 1980s (high confidence)
have been observed. Increasing atmospheric CO2, warming at high
Northern Hemisphere spring snow cover has decreased since at latitudes, and land management interventions have contributed
least 1978 (very high confidence), and there is high confidence to the observed greening trend, but there is low confidence in
that trends in snow cover loss extend back to 1950. It is very likely their relative roles. There is medium confidence that increased
that human influence contributed to these reductions. Earlier onset plant growth associated with CO2 fertilization is the main driver
of snowmelt has contributed to seasonally dependent changes of the observed increase in amplitude of the seasonal cycle of
in streamflow (high confidence). A further decrease of Northern atmospheric CO2 in the Northern Hemisphere. Reactive nitrogen,
Hemisphere seasonal snow cover extent is virtually certain under ozone and aerosols affect terrestrial vegetation and carbon cycle
further global warming (Figure TS.12d). {2.3.2, 3.4.2, 8.3.2. 9.5.3, through deposition and effects on large-scale radiation (high
12.4, 9.2, 11.2, Atlas 8.2} confidence), but the magnitude of these effects on the land carbon
sink, ecosystem productivity and indirect CO2 forcing remains
The frequency and intensity of heavy precipitation events have uncertain. {2.3.4, 3.6.1, 5.2.1, 6.4.5, 12.3.7, 12.4}
increased over a majority of land regions with good observational
coverage since 1950 (high confidence, Box TS.6, Table TS.2). Human Over the last century, there has been a poleward and upslope shift
influence is likely the main driver of this change (Table TS.2). It is in the distribution of many land species (very high confidence)
extremely likely that on most land regions heavy precipitation will as well as increases in species turnover within many ecosystems
TS become more frequent and more intense with additional global (high confidence). There is high confidence that the geographical
warming (Table TS.2, Figure TS.12b). The projected increase in heavy distribution of climate zones has shifted in many parts of the
precipitation extremes translates to an increase in the frequency and world in the last half century. The SRCCL concluded that continued
magnitude of pluvial floods (high confidence) (Table TS.2). {Cross- warming will exacerbate desertification processes (medium
Chapter Box 3.2, 8.4.1, 11.4.2, 11.4.4, 11.5.5, 12.4} confidence) and that ecosystems will become increasingly exposed
to climates beyond those that they are currently adapted to (high
The probability of compound extreme events has likely increased confidence). There is medium confidence that climate change will
due to human-induced climate change. Concurrent heatwaves and increase disturbance by, for example, fire and tree mortality, across
droughts have become more frequent over the last century, and this several ecosystems. Increases are projected in drought, aridity and
trend will continue with higher global warming (high confidence). fire weather in some regions (Section TS.4.3; high confidence).
The probability of compound flooding (storm surge, extreme There is low confidence in the magnitude of these changes, but
rainfall and/or river flow) has increased in some locations and will the probability of crossing uncertain regional thresholds (e.g., fires,
continue to increase due to both sea level rise and increases in heavy forest dieback) increases with further warming (high confidence).
precipitation, including changes in precipitation intensity associated The response of biogeochemical cycles to the anthropogenic
with tropical cyclones (high confidence). {11.8.1, 11.8.2, 11.8.3} perturbation can be abrupt at regional scales, and irreversible on
decadal to century time scales (high confidence). {2.3.4, 5.4.3,
Changes in key aspects of the terrestrial biosphere, such as an 5.4.9, 11.6, 11.8, 12.5, SRCCL 2.2, SRCCL 2.5, SR1.5 3.4}
increase of the growing season length in much of the Northern
84
Technical Summary
Human-caused climate change has driven detectable changes in the global water cycle since the mid-20th century (high
confidence), and it is projected to cause substantial further changes at both global and regional scales (high confidence).
Global land precipitation has likely increased since 1950, with a faster increase since the 1980s (medium confidence).
Atmospheric water vapour has increased throughout the troposphere since at least the 1980s (likely). Annual global
land precipitation will increase over the 21st century as global surface temperature increases (high confidence).
Human influence has been detected in amplified surface salinity and precipitation minus evaporation (P–E) patterns
over the ocean (high confidence).
The severity of very wet and very dry events increase in a warming climate (high confidence), but changes in
atmospheric circulation patterns affect where and how often these extremes occur. Water cycle variability and related
extremes are projected to increase faster than mean changes in most regions of the world and under all emissions
scenarios (high confidence).
Over the 21st century, the total land area subject to drought will increase and droughts will become more frequent
and severe (high confidence). Near-term projected changes in precipitation are uncertain mainly because of internal
variability, model uncertainty and uncertainty in forcings from natural and anthropogenic aerosols (medium
confidence).
Over the 21st century and beyond, abrupt human-caused changes to the water cycle cannot be excluded (medium TS
confidence). {2.3, 3.3, 4.3, 4.4, 4.5, 4.6, 8.2, 8.3, 8.4, 8.5, 8.6, 11.4, 11.6, 11.9}
There is high confidence that the global water cycle has intensified since at least 1980 expressed by, for example, increased atmospheric
moisture fluxes and amplified precipitation minus evaporation patterns. Global land precipitation has likely increased since 1950,
with a faster increase since the 1980s (medium confidence), and a likely human contribution to patterns of change, particularly for
increases in high-latitude precipitation over the Northern Hemisphere. Increases in global mean precipitation are determined by a
robust response to global surface temperature (very likely 2–3% per °C) that is partly offset by fast atmospheric adjustments to
atmospheric heating by greenhouse gases (GHGs) and aerosols (Section TS.3.2.2). The overall effect of anthropogenic aerosols is
to reduce global precipitation through surface radiative cooling effects (high confidence). Over much of the 20th century, opposing
effects of GHGs and aerosols on precipitation have been observed for some regional monsoons (high confidence) (Box TS.13). Global
annual precipitation over land is projected to increase on average by 2.4% (–0.2% to +4.7% likely range) under SSP1-1.9, 4.6% (1.5%
to 8.3% likely range) under SSP2-4.5, and 8.3% (0.9% to 12.9% likely range) under SSP5-8.5 by 2081–2100 relative to 1995–2014
(Box TS.6, Figure 1). Inter-model differences and internal variability contribute to a substantial range in projections of large-scale and
regional water cycle changes (high confidence). The occurrence of volcanic eruptions can alter the water cycle for several years (high
confidence). Projected patterns of precipitation change exhibit substantial regional differences and seasonal contrast as global surface
temperature increases over the 21st century (Box TS.6, Figure 1). {2.3.1, 3.3.2, 3.3.3, 3.5.2, 4.3.1, 4.4.1, 4.5.1, 4.6.1, Cross-Chapter
Box 4.1, 8.2.1, 8.2.2, 8.2.3, Box 8.1, 8.3.2.4, 8.4.1, 8.5.2, 10.4.2}
Global total column water vapour content has very likely increased since the 1980s, and it is likely that human influence has contributed
to tropical upper tropospheric moistening. Near-surface specific humidity has increased over the ocean (likely) and land (very likely)
since at least the 1970s, with a detectable human influence (medium confidence). Human influence has been detected in amplified
surface salinity and precipitation minus evaporation (P–E) patterns over the ocean (high confidence). It is virtually certain that
evaporation will increase over the ocean and very likely that evapotranspiration will increase over land, with regional variations under
future surface warming (Box TS.6, Figure 1). There is high confidence that projected increases in precipitation amount and intensity will
be associated with increased runoff in northern high latitudes (Box TS.6, Figure 1). In response to cryosphere changes (Section TS.2.5),
there have been changes in streamflow seasonality, including an earlier occurrence of peak streamflow in high-latitude and mountain
catchments (high confidence). Projected runoff (Box TS.6, Figure 1c) is typically decreased by contributions from small glaciers because
of glacier mass loss, while runoff from larger glaciers will generally increase with increasing global warming levels until their mass
becomes depleted (high confidence). {2.3.1, 3.3.2, 3.3.3, 3.5.2, 8.2.3, 8.4.1, 11.5}
85
Technical Summary
Warming over land drives an increase in atmospheric evaporative demand and in the severity of drought events (high confidence). Greater
warming over land than over the ocean alters atmospheric circulation patterns and reduces continental near-surface relative humidity,
which contributes to regional drying (high confidence). A very likely decrease in relative humidity has occurred over much of the global
land area since 2000. Projected increases in evapotranspiration due to growing atmospheric water demand will decrease soil moisture
over the Mediterranean region, south-western North America, South Africa, South-Western South America and south-western Australia
(high confidence) (Box TS.6, Figure 1). Some tropical regions are also projected to experience enhanced aridity, including the Amazon
basin and Central America (high confidence). The total land area subject to increasing drought frequency and severity will expand (high
confidence), and in the Mediterranean, South-Western South America, and Western North America, future aridification will far exceed the
magnitude of change seen in the last millennium (high confidence). {4.5.1, 8.2.2, 8.2.3, 8.4.1, Box 8.2, 11.6, 11.9}
Land-use change and water extraction for irrigation have influenced local and regional responses in the water cycle (high confidence).
Large-scale deforestation likely decreases evapotranspiration and precipitation and increases runoff over the deforested regions
relative to the regional effects of climate change (medium confidence). Urbanization increases local precipitation (medium confidence)
and runoff intensity (high confidence) (Box TS.14). Increased precipitation intensities have enhanced groundwater recharge, most
notably in tropical regions (medium confidence). There is high confidence that groundwater depletion has occurred since at least the
start of the 21st century, as a consequence of groundwater withdrawals for irrigation in agricultural areas in drylands. {8.2.3, 8.3.1,
11.1.6, 11.4, 11.6, FAQ 8.1}
Water cycle variability and related extremes are projected to increase faster than mean changes in most regions of the world
TS and under all emissions scenarios (high confidence). A warmer climate increases moisture transport into weather systems, which
intensifies wet seasons and events (high confidence). The magnitudes of projected precipitation increases and related extreme
events depend on model resolution and the representation of convective processes (high confidence). Increases in near-surface
atmospheric moisture capacity of about 7% per 1ºC of warming lead to a similar response in the intensification of heavy precipitation
from sub-daily up to seasonal time scales, increasing the severity of flood hazards (high confidence). The average and maximum
rain-rates associated with tropical and extratropical cyclones, atmospheric rivers and severe convective storms will therefore also
increase with future warming (high confidence). For some regions, there is medium confidence that peak tropical cyclone rain-
rates will increase by more than 7% per 1°C of warming due to increased low-level moisture convergence caused by increases
in wind intensity. In the tropics year-round and in the summer season elsewhere, interannual variability of precipitation and
runoff over land is projected to increase at a faster rate than changes in seasonal mean precipitation (Figure TS.12e,f) (medium
confidence). Sub-seasonal precipitation variability is also projected to increase, with fewer rainy days but increased daily mean
precipitation intensity over many land regions (high confidence). {4.5.3, 8.2.3, 8.4.1, 8.4.2, 8.5.1, 8.5.2, 11.4, 11.5, 11.7, 11.9}
86
Technical Summary
Box TS.6, Figure 1 | Projected water cycle changes. The intent of this figure is to give a geographical overview of changes in multiple components of the global
water cycle using an intermediate emissions scenario. Important key message: without drastic reductions in greenhouse gas emissions, human-induced global warming
will be associated with widespread changes in all components of the water cycle. Long-term (2081–2100) projected annual mean changes (%) relative to present-day
(1995–2014) in the SSP2-4.5 emissions scenario for (a) precipitation, (b) surface evapotranspiration, (c) total runoff and (d) surface soil moisture. Numbers in top
right of each panel indicate indicate the number of Coupled Model Intercomparison Project Phase 6 (CMIP6) models used for estimating the ensemble mean. For other
scenarios, please refer to relevant figures in Chapter 8. Uncertainty is represented using the simple approach: No overlay indicates regions with high model agreement,
where ≥80% of models agree on sign of change; diagonal lines indicate regions with low model agreement, where <80% of models agree on sign of change. For
more information on the simple approach, please refer to the Cross-Chapter Box Atlas.1. {8.4.1; Figures 8.14, 8.17, 8.18, and 8.19}
87
Technical Summary
Climate futures
The climate change that people will experience this century and beyond depends on our
greenhouse gases emissions, how much global warming this will cause and the
response of the climate system to this warming.
Emissions pathways
Different social and economic developments can lead to substantially different future emissions of
carbon dioxide (CO 2 ), other greenhouse gases and air pollutants for the rest of the century.
Today
CO2 peak
140
Very high
CO2 emissions (billion tonnes CO2 per year)
120
TS 60
CO2 peak
40
CO2 halved
20 CO2 peak
CO2 peak Medium
0
Low
Net CO2 removal Very low
-20
2000 2020 2040 2060 2080 2100
0 0.5 1 1.5 2 3 4
Very high
High
Medium
Low
Very low
1980 2000 2020 2040 2060 2080
Today
Infographic TS.1 | Climate Futures. The intent of this figure is to show possible climate futures: The climate change that people will experience this century and beyond
depends on our greenhouse gas emissions, how much global warming this will cause and the response of the climate system to this warming.
(top left) Annual emissions of CO2 for the five core Shared Socio-economic Pathway (SSP) scenarios (very low: SSP1-1.9, low: SSP1-2.6, intermediate: SSP2-4.5, high: SSP3-7.0,
very high: SSP5-8.5).(bottom left) Projected warming for each of these emissions scenarios.
88
Technical Summary
Climate futures
Drought
A drought that used to occur x2.0 x2.4 x4.1
once in a decade now (x1.0 to 5.1) (x1.3 to 5.8) (x1.7 to 7.2)
happens x times more
x1.7
(x0.7 to 4.1)
Precipitation TS
What used to be a wettest x1.5 x1.7 x2.7
day in a decade now (x1.4 to 1.7) (x1.6 to 2.0) (x2.3 to 3.6)
happens x times more x1.3
(x1.2 to1.4)
Snow
Snow cover extent -5% -9% -26%
change (%)
-1% (-13 to 2)
(-3 to 1) (-7 to 2) (-35 to -15)
Tropical cyclones
Proportion of intense tropical +10% +13% +30%
cyclones (%)
The future...
The climate we and the young generations will experience depends on future emissions.
Reducing emissions rapidly will limit further changes, but continued emissions will trigger larger,
faster changes that will increasingly affect all regions. Some changes will persist for hundreds or
thousands of years, so today’s choices will have long-lasting consequences.
(top right) Response of some selected climate variables to four levels of global warming (°C). Changes in the ‘Today’ column are based on a global warming level of 1°C.
(bottom right) The long-term effect of each global warming level on sea level. See Section TS.1.3.1 for more detail on the SSP climate change scenarios.
This infographic builds from several figures in the Technical Summary: Figure TS.4 (for top left panel), Figure TS.6 (bottom left), Figure TS.12 (top right) and Box TS.4, Figure 1b
(bottom right).
89
Technical Summary
TS.3 Understanding the Climate System TS.3.1 Radiative Forcing and Energy Budget
Response and Implications for Limiting
Global Warming Since AR5, the accumulation of energy in the Earth system,
quantified by observations of warming of the ocean,
This section summarizes advances in our knowledge of Earth’s atmosphere, and land and melting of ice, has become
energy budget, including the time evolution of forcings and climate established as a robust measure of the rate of global climate
feedbacks that lead to the climate system responses summarized change on interannual-to-decadal time scales. Compared to
in Section TS.2. It assesses advances since AR5 and SR1.5 in the changes in global surface temperature, the increase in the
estimation of remaining carbon budgets, the Earth system response to global energy inventory exhibits less variability, and thus
carbon dioxide removal, and the quantification of metrics that allow better indicates underlying climate trends.
comparisons of the relative effects of different forcing agents. The
section also highlights: future climate and air pollution responses due The global energy inventory increased by 282 [177 to
to projected changes in short-lived climate forcers (SLCFs); the state 387] zettajoules (ZJ, equal to 1021 Joules) for the period
of understanding of the climate response to potential interventions 1971–2006 and 152 [100 to 205] ZJ for the period 2006–
related to solar radiation modification (SRM); and irreversibility, 2018 (Figure TS.13), with more than 90% accounted for
tipping points and abrupt changes in the climate system.
(a) Global energy inventory (b) Integrated radiative forcing (c) Integrated radiative response
TS
(d) Energy inventory components (e) Radiative forcing components (f ) Energy budget 1971–2018
–
–
2018 2018
Figure TS.13 | Estimates of the net cumulative energy change (ZJ = 1021 Joules) for the period 1971–2018 associated with (a) observations of changes in the global
energy inventory, (b) integrated radiative forcing, and (c) integrated radiative response. The intent is to show assessed changes in energy budget and effective radiative forcings
(ERFs). Black dotted lines indicate the central estimate with likely and very likely ranges as indicated in the legend. The grey dotted lines indicate the energy change associated with an estimated
1850–1900 Earth energy imbalance of 0.2 W m–2 (panel a) and an illustration of an assumed pattern effect of –0.5 W m–2 °C–1 (panel c). Background grey lines indicate equivalent heating
rates in W m–2 per unit area of Earth’s surface. Panels (d) and (e) show the breakdown of components, as indicated in the legend, for the global energy inventory and integrated radiative
forcing, respectively. Panel (f) shows the global energy budget assessed for the period 1971–2018, that is, the consistency between the change in the global energy inventory relative to
1850–1900 and the implied energy change from integrated radiative forcing plus integrated radiative response under a number of different assumptions, as indicated in the figure legend,
including assumptions of correlated and uncorrelated uncertainties in forcing plus response. Shading represents the very likely range for observed energy change relative to 1850–1900 and
likely range for all other quantities. Forcing and response time series are expressed relative to a baseline period of 1850–1900. {Box 7.2, Figure 1}
90
Technical Summary
by ocean warming. To put these numbers in context, the estimates of the global energy imbalance, and closure of the global
2006–2018 average Earth energy imbalance is equivalent sea level budget have led to a strengthened assessment relative
to approximately 20 times the annual rate of global energy to AR5. (high confidence) {7.2.2, 7.5.2.3, Box 7.2, Table 7.1, 9.6.1,
consumption in 2018. The accumulation of energy is driven Cross-Chapter Box 9.1, Table 9.5}
by a positive total anthropogenic effective radiative forcing
(ERF) relative to 1750. As in AR5, the perturbations to Earth’s top-of-atmosphere energy
budget are quantified using ERFs (see also Section TS.2.2). These
The best estimate ERF of 2.72 W m−2 has increased by 0.43 include any consequent adjustments to the climate system (e.g.,
W m–2 relative to that given in AR5 (for 1750–2014) due from changes in atmospheric temperatures, clouds and water vapour
to an increase in the greenhouse gas ERF that is partly as shown in Figure TS.14), but exclude any surface temperature
compensated by a more negative aerosol ERF compared response. Since AR5, ERFs have been estimated for a larger number
to AR5. The greenhouse gas ERF has been revised due to of forcing agents and shown to be more closely related to the
changes in atmospheric concentrations and updates to temperature response than the stratospheric-temperature-adjusted
forcing efficiencies, while the revision to aerosol ERF is due radiative forcing. (high confidence) {7.3.1}
to increased understanding of aerosol–cloud interactions
and is supported by improved agreement between Improved quantifications of ERF, the climate system radiative
different lines of evidence. Improved quantifications response, and the observed energy increase in the Earth system
of ERF, the climate system radiative response, and the for the period 1971–2018 demonstrate improved closure of the
observed energy increase in the Earth system for the global energy budget relative to AR5 (Figure TS.13). Combining
period 1971–2018 demonstrate improved closure of the the likely range of ERF over this period with the central estimate
global energy budget (i.e., the extent to which the sum of radiative response gives an expected energy gain of 340 [47 to
of the integrated forcing and the integrated radiative 662] ZJ. Both estimates are consistent with an independent
response equals the energy gain of the Earth system) observation-based assessment of the global energy increase of TS
compared to AR5 (high confidence). (See FAQ 7.1). {7.2.2, 284 [96 to 471] ZJ (very likely range), expressed relative to the
7.3.5, 7.5.2, Box 7.2, Table 7.1} estimated 1850–1900 Earth energy imbalance. (high confidence)
{7.2.2, 7.3.5, Box 7.2}
The global energy inventory change for the period 1971–2006
corresponds to an Earth energy imbalance (Box TS.1) of 0.50 [0.32 The assessed greenhouse gas ERF over the 1750–2019 period
to 0.69] W m–2, increasing to 0.79 [0.52 to 1.06] W m–2 for the period (Section TS.2.2) has increased by +0.59 W m−2 over AR5 estimates
2006–2018. Ocean heat uptake is by far the largest contribution and for 1750–2011. This increase includes +0.34 W m–2 from increases in
accounts for 91% of the total energy change. Land warming, melting atmospheric concentrations of well-mixed greenhouse gases (including
of ice and warming of the atmosphere account for about 5%, 3% and halogenated species) since 2011, +0.15 W m –2 from upwards revisions
1% of the total change, respectively. More comprehensive analysis of of their radiative efficiencies and +0.10 W m –2 from re-evaluation of
inventory components, cross-validation of satellite and in situ-based the ozone and stratospheric water vapour ERF. {7.3.2, 7.3.4, 7.3.5}
Absorbed Cloud
sunlight feedback
Adjustments
in clouds, water
vapour, temperature
fewer low clouds
Longwave Increased greenhouse Increased greenhouse Increased greenhouse
radiative gases or aerosols gases or aerosols gases or aerosols
emission Surface and Surface albedo and
vegetation biogeochemical
response feedback
Increased
water
vapour
No change in meteorology or surface No change in surface temperature Increased surface temperature
temperature
Figure TS.14 | Schematic representation of changes in the top-of-atmosphere (TOA) radiation budget following a perturbation. The intent of this figure is to
illustrate the concept of adjustments in the climate system following a perturbation in the radiation budget. The baseline TOA energy budget (a) responds instantaneously to
perturbations (b), leading to adjustments in the atmospheric meteorology and composition and land surface that are independent of changes in surface temperature (c). Surface
temperature changes (here using an increase as an example) lead to physical, biogeophysical and biogeochemical feedback processes (d). Long-term feedback processes, such
as those involving ice sheets, are not shown here. {adapted from Figure 7.2; FAQ 7.2, Figure 1; and Figure 8.3}
91
Technical Summary
TS
Figure TS.15 | Contribution to (a) effective radiative forcing (ERF) and (b) global surface temperature change from component emissions for 1750–2019
based on Coupled Model Intercomparison Project Phase 6 (CMIP6) models and (c) net aerosol ERF for 1750–2014 from different lines of evidence. The
intent of this figure is to show advances since AR5 in the understanding of (a) emissions-based ERF, (b) global surface temperature response for short-lived climate forcers as
estimated in Chapter 6, and (c) aerosol ERF from different lines of evidence as assessed in Chapter 7. In panel (a), ERFs for well-mixed greenhouse gases (WMGHGs) are
from the analytical formulae. ERFs for other components are multi-model means based on Earth system model simulations that quantify the effect of individual components.
The derived emissions-based ERFs are rescaled to match the concentration-based ERFs in Figure 7.6. Error bars are 5–95% and for the ERF account for uncertainty in radiative
efficiencies and multi-model error in the means. In panel (b), the global mean temperature response is calculated from the ERF time series using an impulse response function.
In panel (c), the AR6 assessment is based on energy balance constraints, observational evidence from satellite retrievals, and climate model-based evidence. For each line of
evidence, the assessed best-estimate contributions from ERF due to aerosol–radiation interactions (ERFari) and aerosol–cloud interactions (ERFaci) are shown with darker and
paler shading, respectively. Estimates from individual CMIP Phase 5 (CMIP5) and CMIP6 models are depicted by blue and red crosses, respectively. The observational assessment
for ERFari is taken from the instantaneous forcing due to aerosol–radiation interactions (IRFari). Uncertainty ranges are given in black bars for the total aerosol ERF and depict
very likely ranges. {6.4.2, Figure 6.12, 7.3.3, Cross-Chapter Box 7.1, Table 7.8, Figure 7.5}
For CO2, CH4, N2O, and chlorofluorocarbons, there is now evidence to emissions make the dominant contribution to the ERF from aerosol–
quantify the effect on ERF of tropospheric adjustments. The assessed cloud interactions (high confidence). Over the 1750–2019 period,
ERF for a doubling of CO2 compared to 1750 levels (3.9 ± 0.5 Wm–2) is the contributions from the emitted compounds to global surface
larger than in AR5. For CO2, the adjustments include the physiological temperature changes broadly match their contributions to the ERF
effects on vegetation. The reactive well-mixed greenhouse gases (high confidence) (Figure TS.15b). Since a peak in emissions-induced
(CH4, N2O, and halocarbons) cause additional chemical adjustments SO2 ERF has already occurred recently (Section TS.2.2) and since there
to the atmosphere through changes in ozone and aerosols (Figure is a delay in the full global surface temperature response owing to
TS.15a). The ERF due to CH4 emissions is 1.19 [0.81 to 1.58] W m–2, of the thermal inertia in the climate system, changes in SO2 emissions
which 0.35 [0.16 to 0.54] W m–2 is attributed to chemical adjustments have a slightly larger contribution to global surface temperature
mainly via ozone. These chemical adjustments also affect the change compared with changes in CO2 emissions, relative to their
emissions metrics (Section TS.3.3.3). Changes in sulphur dioxide (SO2) respective contributions to ERF. {6.4.2, 7.3.2}
92
Technical Summary
Aerosols contributed an ERF of –1.3 [–2.0 to –0.6] W m–2 over the evolves and global surface temperature increases, leading to an ECS
period 1750 to 2014 (medium confidence). The ERF due to aerosol– that is higher than was inferred in AR5 based on warming over the
cloud interactions (ERFaci) contributes most to the magnitude of instrumental record (high confidence). Historical surface temperature
the total aerosol ERF (high confidence) and is assessed to be –1.0 change since 1870 has shown relatively little warming in several key
[–1.7 to –0.3] W m–2 (medium confidence), with the remainder due regions of positive feedbacks, including the eastern equatorial Pacific
to aerosol–radiation interactions (ERFari), assessed to be –0.3 [–0.6 Ocean and the Southern Ocean, while showing greater warming
to 0.0] W m–2 (medium confidence). There has been an increase in in key regions of negative feedbacks, including the western Pacific
the estimated magnitude – but a reduction in the uncertainty – of warm pool. Based on process understanding, climate modelling, and
the total aerosol ERF relative to AR5, supported by a combination paleoclimate reconstructions of past warm periods, it is expected
of increased process-understanding and progress in modelling and that future warming will become enhanced over the eastern Pacific
observational analyses (Figure TS.15c). Effective radiative forcing Ocean (medium confidence) and Southern Ocean (high confidence)
estimates from these separate lines of evidence are now consistent on centennial time scales. This new understanding, along with
with each other, in contrast to AR5, and support the assessment that updated estimates of historical temperature change, ERF, and energy
it is virtually certain that the total aerosol ERF is negative. Compared imbalance, reconciles previously disparate ECS estimates. {7.4.4,
to AR5, the assessed magnitude of ERFaci has increased, while that 7.5.2, 7.5.3}
of ERFari has decreased. {7.3.3, 7.3.5}
The AR6 best estimate of ECS is 3°C, the likely range is 2.5°C to
4°C and the very likely range is 2°C to 5°C. There is a high level
TS.3.2 Climate Sensitivity and Earth System Feedbacks of agreement among the four main lines of evidence listed above
(Figure TS.16b), and altogether it is virtually certain that ECS is larger
TS.3.2.1 Equilibrium Climate Sensitivity, Transient Climate than 1.5°C, but currently it is not possible to rule out ECS values
Response, and Transient Climate Response to above 5°C. Therefore, the 5°C upper end of the very likely range is
Cumulative Carbon-dioxide Emissions assessed with medium confidence and the other bounds with high TS
confidence. {7.5.5}
Since AR5, substantial quantitative progress has been
made in combining new evidence of Earth’s climate Based on process understanding, warming over the instrumental
sensitivity with improvements in the understanding and record, and emergent constraints, the best estimate of TCR is 1.8°C,
quantification of Earth’s energy imbalance, the instrumental the likely range is 1.4°C to 2.2°C and the very likely range is 1.2°C to
record of global surface temperature change, paleoclimate 2.4°C. There is a high level of agreement among the different lines of
change from proxy records, climate feedbacks and their evidence (Figure TS.16c) (high confidence). {7.5.5}
dependence on time scale and climate state. A key advance
is the broad agreement across these multiple lines of On average, CMIP6 models have higher mean ECS and TCR values
evidence, supporting a best estimate of equilibrium climate than the CMIP5 generation of models and also have higher mean
sensitivity of 3°C, with a very likely range of 2°C to 5°C. The values and wider spreads than the assessed best estimates and very
likely range of 2.5°C to 4°C is narrower than the AR5 likely likely ranges within this Report. These higher mean ECS and TCR
range of 1.5°C to 4.5°C. {7.4, 7.5} values can be traced to a positive net cloud feedback that is larger in
CMIP6 by about 20%. The broader ECS and TCR ranges from CMIP6
Constraints on equilibrium climate sensitivity (ECS) and transient also lead the models to project a range of future warming that
climate response (TCR) (see Glossary) are based on four main lines is wider than the assessed future warming range, which is based
of evidence: feedback process understanding, climate change and on multiple lines of evidence (Cross-Section Box TS.1). However,
variability seen within the instrumental record, paleoclimate evidence, some of the high-sensitivity CMIP6 models (Section TS.1.2.2) are
and so-called ‘emergent constraints’, whereby a relationship less consistent with observed recent changes in global warming
between an observable quantity and either ECS or TCR established and with paleoclimate proxy records than models with ECS within
within an ensemble of models is combined with observations to the very likely range. Similarly, some of the low-sensitivity models
derive a constraint on ECS or TCR. In reports up to and including are less consistent with the paleoclimate data. The CMIP6 models
the IPCC Third Assessment Report, ECS and TCR derived directly with the highest ECS and TCRs values provide insights into low-
from ESMs were the primary line of evidence. However, since AR4, likelihood, high-impact futures, which cannot be excluded based on
historical warming and paleoclimates provided useful additional currently available evidence (Cross-Section Box TS.1). {4.3.1, 4.3.4,
evidence (Figure TS.16a). This Report differs from previous reports in 7.4.2, 7.5.6}
not directly using climate model estimates of ECS and TCR in the
assessed ranges of climate sensitivity. {1.5, 7.5} Uncertainties regarding the true value of ECS and TCR are the dominant
source of uncertainty in global temperature projections over the
It is now clear that when estimating ECS and TCR, the dependence 21st century under moderate to high GHG concentrations scenarios.
of feedbacks on time scales and the climate state must be accounted For scenarios that reach net zero CO2 emissions (Section TS.3.3), the
for. Feedback processes are expected to become more positive uncertainty in the ERF values of aerosol and other SLCFs contribute
overall (more amplifying of global surface temperature changes) on substantial uncertainty in projected temperature. Global ocean heat
multi-decadal time scales as the spatial pattern of surface warming uptake is a smaller source of uncertainty in centennial warming. {7.5.7}
93
Technical Summary
SAR
AR4
AR5
FAR
Charney
TAR
4 Likely: 2.5–4°C 4
AR6
3 Best estimate: 3°C 3
2 2
AR6 combines evidence from:
· Process understanding
· Instrumental record
1 Primarily model evidence p < 5%
·
· Paleoclimates
1
Emergent constraints
Also considers instrumental record and paleoclimates
(b) Equilibrium climate sensitivity (ºC) assessed in AR6 and (c) Transient climate response (ºC) assessed in AR6 and
simulated by CMIP6 ESMs simulated by CMIP6 ESMs
Paleoclimates Paleoclimates
0 1 2 3 4 5 6 7 8 9 0 1 2 3 4
Best estimate range or value Likely range or limit Very likely range or limit Extremely likely limit
Figure TS.16 | (a) Evolution of equilibrium climate sensitivity (ECS) assessments from the Charney Report through a succession of IPCC Assessment
Reports to AR6, and lines of evidence and combined assessment for (b) ECS and (c) transient climate response (TCR) in AR6. The intent of this figure is to
show the progression in estimates of ECS, including uncertainty and the lines of evidence used for assessment, and to show the lines of assessment used to assess ECS and
TCR in AR6. In panel (a), the lines of evidence considered are listed below each assessment. Best estimates are marked by horizontal bars, likely ranges by vertical bars, and
very likely ranges by dotted vertical bars. In panel (b) and (c), assessed ranges are taken from Tables 7.13 and 7.14 for ECS and TCR respectively. Note that for the ECS
assessment based on both the instrumental record and paleoclimates, limits (i.e., one-sided distributions) are given, which have twice the probability of being outside the
maximum/minimum value at a given end, compared to ranges (i.e., two tailed distributions) which are given for the other lines of evidence. For example, the extremely likely limit
of greater than 95% probability corresponds to one side of the very likely (5% to 95%) range. Best estimates are given as either a single number or by a range represented by
grey box. Coupled Model Intercomparison Project Phase 6 (CMIP6) Earth system model (ESM) values are not directly used as a line of evidence but are presented on the figure
for comparison. {1.5, 7.5; Tables 7.13 and 7.14; Figure 7.18}
The transient climate response to cumulative CO2 emissions (TCRE) of TCR. Beyond this century, there is low confidence that the TCRE
is the ratio between globally averaged surface temperature increase alone remains an accurate predictor of temperature changes in
and cumulative CO2 emissions (see Glossary). This Report reaffirms scenarios of very low or net negative CO2 emissions because of
with high confidence the finding of AR5 that there is a near-linear uncertain Earth system feedbacks that can result in further changes
relationship between cumulative CO2 emissions and the increase in in temperature or a path dependency of warming as a function of
global average temperature caused by CO2 over the course of this cumulative CO2 emissions. {4.6.2, 5.4, 5.5.1}
century for global warming levels up to at least 2°C relative to 1850–
1900. The TCRE falls likely in the 1.0°C–2.3°C per 1000 PgC range, TS.3.2.2 Earth System Feedbacks
with a best estimate of 1.65°C per 1000 PgC. This is equivalent to a
0.27°C–0.63°C range with a best estimate of 0.45°C when expressed The combined effect of all climate feedback processes is to amplify
in units per 1000 GtCO2. This range is about 15% narrower than the the climate response to forcing (virtually certain). While major
0.8°–2.5°C per 1000 PgC assessment of AR5 because of a better advances in the understanding of cloud processes have increased
integration of evidence across chapters, in particular the assessment the level of confidence and decreased the uncertainty range for the
94
Technical Summary
cloud feedback by about 50% compared to AR5, clouds remain the Natural sources and sinks of non-CO2 greenhouse gases such as
largest contribution to overall uncertainty in climate feedbacks (high methane (CH4) and nitrous oxide (N2O) respond both directly and
confidence). Uncertainties in the ECS and other climate sensitivity indirectly to atmospheric CO2 concentration and climate change,
metrics, such as the TCR and TCRE, are the dominant source of and thereby give rise to additional biogeochemical feedbacks
uncertainty in global temperature projections over the 21st century in the climate system. Many of these feedbacks are only partially
under moderate to high GHG emissions scenarios. CMIP6 models understood and are not yet fully included in ESMs. There is medium
have higher mean values and wider spreads in ECS and TCR than the confidence that the net response of natural ocean and land CH4 and
assessed best estimates and very likely ranges within this Report, N2O sources to future warming will be increased emissions, but the
leading the models to project a range of future warming that is magnitude and timing of the responses of each individual process is
wider than the assessed future warming range (Section TS.2.2). {7.1, known with low confidence. {5.4.7}
7.4.2, 7.5}
Non-CO2 biogeochemical feedbacks induced from changes in
Earth system feedbacks can be categorized into three broad groups: emissions, abundances or lifetimes of SLCFs mediated by natural
physical feedbacks, biogeophysical and biogeochemical feedbacks, processes or atmospheric chemistry are assessed to decrease
and feedbacks associated with ice sheets. In previous assessments, ECS (Figure TS.17b). These non-CO2 biogeochemical feedbacks
the ECS has been associated with a distinct set of physical feedbacks are estimated from ESMs, which since AR5 have advanced to
(Planck response, water vapour, lapse rate, surface albedo, and cloud include a consistent representation of biogeochemical cycles and
feedbacks). In this assessment, a more general definition of ECS is atmospheric chemistry. However, process-level understanding
adopted whereby all biogeophysical and biogeochemical feedbacks of many biogeochemical feedbacks involving SLCFs, particularly
that do not affect the atmospheric concentration of CO2 are included. natural emissions, is still emerging, resulting in low confidence in
These include changes in natural CH4 emissions, natural aerosol the magnitude and sign of the feedbacks. The central estimate of
emissions, N2O, ozone, and vegetation, which all act on time scales of the total biogeophysical and non-CO2 biogeochemical feedback is
years to decades and are therefore relevant for temperature change assessed to be −0.01 [–0.27 to +0.25] W m–2 °C–1 (Figure TS.17a). TS
over the 21st century. Because the total biogeophysical and non-CO2 {5.4.7, 5.4.8, 6.2.2, 6.4.5, 7.4, Table 7.10}
biogeochemical feedback is assessed to have a central value that is
near zero (low confidence), including it does not affect the assessed The combined effect of all known radiative feedbacks (physical,
ECS but does contribute to the net feedback uncertainty. The biogeophysical, and non-CO2 biogeochemical) is to amplify the
biogeochemical feedbacks that affect the atmospheric concentration base climate response (in the absence of feedbacks), also known
of CO2 are not included because ECS is defined as the response to as the Planck temperature response20 (virtually certain). Combining
a sustained doubling of CO2. Moreover, the long-term feedbacks these feedbacks with the Planck response, the net climate feedback
associated with ice sheets are not included in the ECS owing to their parameter is assessed to be –1.16 [–1.81 to –0.51] W m–2 °C–1,
long time scales of adjustment. {5.4, 6.4, 7.4, 7.5, Box 7.1} which is slightly less negative than that inferred from the overall ECS
assessment. The combined water vapour and lapse rate feedback
The net effect of changes in clouds in response to global warming is makes the largest single contribution to global warming, whereas
to amplify human-induced warming, that is, the net cloud feedback the cloud feedback remains the largest contribution to overall
is positive (high confidence). Compared to AR5, major advances uncertainty. Due to the state-dependence of feedbacks, as evidenced
in the understanding of cloud processes have increased the level from paleoclimate observations and from models, the net feedback
of confidence and decreased the uncertainty range in the cloud parameter will increase (become less negative) as global temperature
feedback by about 50% (Figure TS.17a). An assessment of the low- increases. Furthermore, on long time scales the ice-sheet feedback
altitude cloud feedback over the subtropical ocean, which was parameter is very likely positive, promoting additional warming on
previously the major source of uncertainty in the net cloud feedback, millennial time scales as ice sheets come into equilibrium with the
is improved owing to a combined use of climate model simulations, forcing. (high confidence) {7.4.2, 7.4.3, Figure 7.14, Table 7.10}
satellite observations, and explicit simulations of clouds, altogether
leading to strong evidence that this type of cloud amplifies global The carbon cycle provides for additional feedbacks on climate owing
warming. The net cloud feedback is assessed to be +0.42 [–0.10 to the sensitivity of land–atmosphere and ocean–atmosphere carbon
to 0.94] W m –2 °C–1. A net negative cloud feedback is very unlikely. fluxes and storage to changes in climate and in atmospheric CO2
The CMIP5 and CMIP6 ranges of cloud feedback are similar to this (Figure TS.17c). Because of the time scales associated with land and
assessed range, with CMIP6 having a slightly more positive median ocean carbon uptake, these feedbacks are known to be scenario
cloud feedback (high confidence). The surface albedo feedback and dependent. Feedback estimates deviate from linearity in scenarios
combined water vapour-lapse rate feedback are positive (Figure of stabilizing or reducing concentrations. With high confidence,
TS.17a), with high confidence in the estimated value of each based increased atmospheric CO2 will lead to increased land and ocean
on multiple lines of evidence, including observations, models and carbon uptake, acting as a negative feedback on climate change. It
theory (Box TS.6). {7.4.2, Figure 7.14, Table 7.10} is likely that a warmer climate will lead to reduced land and ocean
carbon uptake, acting as a positive feedback (Box TS.5). {4.3.2,
5.4.1–5}
20 For reference, the Planck temperature response for a doubling of atmospheric CO2 is approximately 1.2°C at equilibrium.
95
Technical Summary
(b) Biogeophysical and non-CO2 biogeochemical climate feedbacks Mean [5−95% range]
−3.5 −3.0 −2.5 −2.0 −1.5 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Climate feedback parameter (Wm-2 °C-1)
Figure TS.17 | An overview of physical and biogeochemical feedbacks in the climate system. The intent of this figure is to summarize assessed estimates of physical,
biogeophysical and biogeochemical feedbacks on global temperature based on Chapters 5, 6 and 7. (a) Synthesis of physical, biogeophysical and non-carbon dioxide (CO2)
biogeochemical feedbacks that are included in the definition of equilibrium climate sensitivity (ECS) assessed in this Technical Summary. These feedbacks have been assessed
using multiple lines of evidence including observations, models and theory. The net feedback is the sum of the Planck response, water vapour and lapse rate, surface albedo,
cloud, and biogeophysical and non-CO2 biogeochemical feedbacks. Bars denote the mean feedback values, and uncertainties represent very likely ranges; (b) Estimated values of
individual biogeophysical and non-CO2 biogeochemical feedbacks. The atmospheric methane (CH4) lifetime and other non-CO2 biogeochemical feedbacks have been calculated
using global Earth system model simulations from AerChemMIP, while the CH4 and nitrous oxide (N2O) source responses to climate have been assessed for the year 2100 using
a range of modelling approaches using simplified radiative forcing equations. The estimates represent the mean and 5–95% range. The level of confidence in these estimates
is low owing to the large model spread. (c) Carbon-cycle feedbacks as simulated by models participating in the C4MIP of the Coupled Model Intercomparison Project Phase 6
(CMIP6). An independent estimate of the additional positive carbon-cycle climate feedbacks from permafrost thaw, which is not considered in most C4MIP models, is added.
The estimates represent the mean and 5–95% range. Note that these feedbacks act through modifying the atmospheric concentration of CO2 and thus are not included in the
definition of ECS, which assumes a doubling of CO2, but are included in the definition and assessed range of the transient climate response to cumulative CO2 emissions (TCRE).
{5.4.7, 5.4.8, Box 5.1, Figure 5.29, 6.4.5, Table 6.9, 7.4.2, Table 7.10}
96
Technical Summary
Thawing terrestrial permafrost will lead to carbon release (high TS.3.3 Temperature Stabilization, Net Zero Emissions
confidence), but there is low confidence in the timing, magnitude and Mitigation
and the relative roles of CO2 versus CH4 as feedback processes. An
ensemble of models projects CO2 release from permafrost to be TS.3.3.1 Remaining Carbon Budgets and Temperature
3–41 PgC per 1ºC of global warming by 2100, leading to warming Stabilization
strong enough that it must be included in estimates of the remaining
carbon budget but weaker than the warming from fossil fuel burning. The near-linear relationship between cumulative CO2
However, the incomplete representation of important processes, emissions and maximum global surface temperature increase
such as abrupt thaw, combined with weak observational constraints, caused by CO2 implies that stabilizing human-induced global
only allow low confidence in both the magnitude of these estimates temperature increase at any level requires net anthropogenic
and in how linearly proportional this feedback is to the amount of CO2 emissions to become zero. This near-linear relationship
global warming. There is emerging evidence that permafrost thaw further implies that mitigation requirements for limiting
and thermokarst give rise to increased CH4 and N2O emissions, which warming to specific levels can be quantified in terms of a
leads to the combined radiative forcing from permafrost thaw being carbon budget (high confidence). Remaining carbon budget
larger than from CO2 emissions only. However, the quantitative estimates have been updated since AR5 with methodological
understanding of these additional feedbacks is low, particularly for improvements, resulting in larger estimates that are
N2O. These feedbacks, as well as potential additional carbon losses consistent with SR1.5. Several factors, including estimates
due to climate-induced fire feedback are not routinely included in of historical warming, future emissions from thawing
Earth system models. {Box 5.1, 5.4.3, 5.4.7, 5.4.8} permafrost, variations in projected non-CO2 warming, and
the global surface temperature change after cessation of
CO2 emissions, affect the exact value of carbon budgets (high
confidence). {1.3.5, Box 1.2, 4.7.1, 5.5}
TS
(a) Cumulative carbon emissions since 1850 (PgC)
(b)
0 200 400 600 800 1000 1200 1400 1600 1800 2000
7
Global mean temperature increase since 1850–1900 (°C)
SSP3-7.0
SSP5-8.5
remaining
1
historical
0 human-induced
warming
0 (Section 5.5.2.2.2)
-1
0 1000 2000 3000 4000 5000 6000 7000 0 remaining unrepresented Earth system
carbon budget feedbacks (Section 5.5.2.2.5)
Cumulative carbon dioxide emissions since 1850 (GtCO2)
Cumulative CO2 emissions from today (GtCO2)
Figure TS.18 | Illustration of (a) relationship between cumulative emissions of carbon dioxide (CO2) and global mean surface air temperature increase
and (b) the assessment of the remaining carbon budget from its constituting components based on multiple lines of evidence. The intent of this figure is to
show (i) the proportionality between cumulative CO2 emissions and global surface air temperature in observations and models as well as the assessed range of the transient
climate response to cumulative CO2 emissions (TCRE), and (ii) how information is combined to derive remaining carbon budgets consistent with limiting warming to a specific
level. Carbon budgets consistent with various levels of additional warming are provided in Table 5.8 and should not be read from the illustrations in either panel. In panel (a)
thin black line shows historical CO2 emissions together with the assessed global surface temperature increase from 1850–1900 as assessed in Chapter 2 (Box 2.3). The orange-
brown range with its central line shows the estimated human-induced share of historical warming. The vertical orange-brown line shows the assessed range of historical human-
induced warming for the 2010–2019 period relative to 1850–1900 (Chapter 3). The grey cone shows the assessed likely range for the TCRE (Section 5.5.1.4), starting from
2015. Thin coloured lines show Coupled Model Intercomparison Project Phase 6 (CMIP6) simulations for the five scenarios of the WGI core set (SSP1-1.9, light blue; SSP1-2.6,
blue; SSP2-4.5, yellow; SSP3-7.0, red; SSP5-8.5, maroon), starting from 2015 and until 2100. Diagnosed carbon emissions are complemented with estimated land-use change
emissions for each respective scenario. Coloured areas show the Chapter 4 assessed very likely range of global surface temperature projections and thick coloured central lines
show the median estimate, for each respective scenario. These temperature projections are expressed relative to cumulative CO2 emissions that are available for emissions-
driven CMIP6 ScenarioMIP experiments for each respective scenario. For panel (b), the remaining allowable warming is estimated by combining the global warming limit
of interest with the assessed historical human-induced warming (Section 5.5.2.2.2), the assessed future potential non-CO2 warming contribution (Section 5.5.2.2.3) and the
zero emissions commitment (ZEC; Section 5.5.2.2.4). The remaining allowable warming (vertical blue bar) is subsequently combined with the assessed TCRE (Sections 5.5.1.4
and 5.5.2.2.1) and contribution of unrepresented Earth system feedbacks (Section 5.5.2.2.5) to provide an assessed estimate of the remaining carbon budget (horizontal blue
bar, Table 5.8). Note that contributions in panel (b) are illustrative and are not to scale. For example, the central ZEC estimate was assessed to be zero. {Box 2.3, 5.2.1, 5.2.2,
Figure 5.31}
97
Technical Summary
Limiting further climate change would require substantial and but they may vary by an estimated ±220 GtCO2 depending on
sustained reductions of GHG emissions. Without net zero CO2 how deeply future non-CO2 emissions are assumed to be reduced
emissions, and a decrease in the net non-CO2 forcing (or sufficient (Table TS.3). {5.5.2, 5.6, Box 5.2, 7.6}
net negative CO2 emissions to offset any further warming from
net non-CO2 forcing), the climate system will continue to warm. There is high confidence that several factors, including estimates of
There is high confidence that mitigation requirements for limiting historical warming, future emissions from thawing permafrost, and
warming to specific levels over this century can be estimated using variations in projected non-CO2 warming, affect the value of carbon
a carbon budget that relates cumulative CO2 emissions to global budgets but do not change the conclusion that global CO2 emissions
mean temperature increase (Figure TS.18, Table TS.3). For the period would need to decline to net zero to halt global warming. Estimates
1850–2019, a total of 2390 ± 240 GtCO2 of anthropogenic CO2 has may vary by ±220 GtCO2 depending on the level of non-CO2 emissions
been emitted. Remaining carbon budgets (starting from 1 January at the time global anthropogenic CO2 emissions reach net zero levels.
2020) for limiting warming to 1.5°C, 1.7°C and 2.0°C are estimated This variation is referred to as non-CO2 scenario uncertainty and
at 500 GtCO2, 850 GtCO2 and 1350 GtCO2, respectively, based on will be further assessed in the AR6 Working Group III Contribution.
the 50th percentile of TCRE. For the 67th percentile, the respective Geophysical uncertainties surrounding the climate response to these
values are 400 GtCO2, 700 GtCO2 and 1150 GtCO2. The remaining non-CO2 emissions result in an additional uncertainty of at least
carbon budget estimates for different temperature limits assume ±220 GtCO2, and uncertainties in the level of historical warming
that non-CO2 emissions are mitigated consistent with the median result in a ±550 GtCO2 uncertainty. {5.4, 5.5.2}
reductions found in scenarios in the literature as assessed in SR1.5,
Table TS.3 | Estimates of remaining carbon budgets and their uncertainties. Assessed estimates are provided for additional human-induced warming, expressed as
global surface temperature, since the recent past (2010–2019), likely amounted to 0.8° to 1.3°C with a best estimate of 1.07°C relative to 1850–1900. Historical CO2 emissions
between 1850 and 2014 have been estimated at about 2180 ± 240 GtCO2 (1-sigma range), while since 1 January 2015, an additional 210 GtCO2 has been emitted until the
TS end of 2019. GtCO2 values to the nearest 50. {Table 3.1, 5.5.1, 5.5.2, Box 5.2, Table 5.1, Table 5.7, Table 5.8}
17th 33rd 50th 67th 83rd GtCO2 GtCO2 GtCO2 GtCO2 GtCO2
a
Human-induced global surface temperature increase between 1850–1900 and 2010–2019 is assessed at 0.8–1.3°C (likely range; Cross-Section Box TS.1) with a best estimate
of 1.07°C. Combined with a central estimate of TCRE (1.65°C per 1000 PgC) this uncertainty in isolation results in a potential variation of remaining carbon budgets of ±550
GtCO2, which, however, is not independent of the assessed uncertainty of TCRE and thus not fully additional.
b
TCRE: transient climate response to cumulative emissions of carbon dioxide, assessed to fall likely between 1.0–2.3°C per 1000 PgC with a normal distribution, from which
the percentiles are taken. Additional Earth system feedbacks are included in the remaining carbon budget estimates as discussed in Section 5.5.2.2.5.
c
Estimates assume that non-CO2 emissions are mitigated consistent with the median reductions found in scenarios in the literature as assessed in SR1.5. Non-CO2 scenario
variations indicate how much remaining carbon budget estimates vary due to different scenario assumptions related to the future evolution of non-CO2 emissions in mitigation
scenarios from SR1.5 that reach net zero CO2 emissions. This variation is additional to the uncertainty in TCRE. The Working Group III Contribution to AR6 will reassess the
potential for non-CO2 mitigation based on literature since SR1.5.
d
Geophysical uncertainties reported in these columns and TCRE uncertainty are not statistically independent, as uncertainty in TCRE depends on uncertainty in the assessment
of historical temperature, non-CO2 versus CO2 forcing, and uncertainty in emissions estimates. These estimates cannot be formally combined, and these uncertainty variations
are not directly additional to the spread of remaining carbon budgets due to TCRE uncertainty reported in columns three to seven.
e
Recent emissions uncertainty reflects the ±10% uncertainty in the historical CO2 emissions estimate since 1 January 2015.
98
Technical Summary
Methodological improvements and new evidence result in updated to be removed to compensate for an emission of a given
remaining carbon budget estimates. The assessment in AR6 applies magnitude to attain the same change in atmospheric CO2
the same methodological improvements as in SR1.5, which uses (medium confidence). CDR methods have wide-ranging side-
a recent observed baseline for historic temperature change and effects that can either weaken or strengthen the carbon
cumulative emissions. Changes compared to SR1.5 are therefore sequestration and cooling potential of these methods and
small: the assessment of new evidence results in updated median affect the achievement of sustainable development goals
remaining carbon budget estimates for limiting warming to 1.5°C (high confidence). {4.6.3, 5.6}
and 2°C being the same and about 60 GtCO2 smaller, respectively,
after accounting for emissions since SR1.5. Meanwhile, remaining Carbon dioxide removal (CDR) refers to anthropogenic activities
carbon budgets for limiting warming to 1.5°C would be about that deliberately remove CO2 from the atmosphere and durably
300–350 GtCO2 larger if evidence and methods available at the store it in geological, terrestrial or ocean reservoirs, or in products.
time of AR5 would be used. If a specific remaining carbon budget Carbon dioxide is removed from the atmosphere by enhancing
is exceeded, this results in a lower probability of keeping warming biological or geochemical carbon sinks or by direct capture of CO2
below a specified temperature level and higher irreversible global from air. Emissions pathways that limit global warming to 1.5°C or
warming over decades to centuries, or alternatively a need for net 2°C typically assume the use of CDR approaches in combination
negative CO2 emissions or further reductions in non-CO2 greenhouse with GHG emissions reductions. CDR approaches could be used to
gases after net zero CO2 is achieved to return warming to lower levels compensate for residual emissions from sectors that are difficult or
in the long term. {5.5.2, 5.6, Box 5.2} costly to decarbonize. CDR could also be implemented at a large scale
to generate global net negative CO2 emissions (i.e., anthropogenic
Based on idealized model simulations that explore the climate CO2 removals exceeding anthropogenic emissions), which could
response once CO2 emissions have been brought to zero, the compensate for earlier emissions as a way to meet long-term climate
magnitude of the zero CO2 emissions commitment (ZEC, see Glossary) stabilization goals after a temperature overshoot. This Report assesses
is assessed to be likely smaller than 0.3°C for time scales of about the effects of CDR on the carbon cycle and climate. Co-benefits and TS
half a century and cumulative CO2 emissions broadly consistent with trade-offs for biodiversity, water and food production are briefly
global warming of 2°C. However, there is low confidence about its discussed for completeness, but a comprehensive assessment of the
sign on time scales of about half a century. For lower cumulative CO2 ecological and socio-economic dimensions of CDR options is left to
emissions, the range would be smaller yet with equal uncertainty the WGII and WGIII reports. {4.6.3, 5.6}
about the sign. If the ZEC is positive on decadal time scales, additional
warming leads to a reduction in the estimates of remaining carbon CDR methods have the potential to sequester CO2 from the
budgets, and vice versa if it is negative. {4.7.1, 5.5.2} atmosphere (high confidence). In the same way part of current
anthropogenic net CO2 emissions are taken up by land and ocean
Permafrost thaw is included in estimates together with other carbon stores, net CO2 removal will be partially counteracted by CO2
feedbacks that are often not captured by models. Limitations in release from these stores, such that the amount of CO2 sequestered
modelling studies combined with weak observational constraints by CDR will not result in an equivalent drop in atmospheric CO2 (very
only allow low confidence in the magnitude of these estimates high confidence). The fraction of CO2 removed from the atmosphere
(Section TS.3.2.2). Despite the large uncertainties surrounding the that is not replaced by CO2 released from carbon stores – a measure
quantification of the effect of additional Earth system feedback of CDR effectiveness – decreases slightly with increasing amounts of
processes, such as emissions from wetlands and permafrost thaw, removal (medium confidence) and decreases strongly if CDR is applied
these feedbacks represent identified additional risk factors that at lower atmospheric CO2 concentrations (medium confidence). The
scale with additional warming and mostly increase the challenge of reduction in global surface temperature is approximately linearly
limiting warming to specific temperature levels. These uncertainties related to cumulative CO2 removal (high confidence). Because of this
do not change the basic conclusion that global CO2 emissions would near-linear relationship, the amount of cooling per unit CO2 removed
need to decline to net zero to halt global warming. {5.4.8, 5.5.2, is approximately independent of the rate and amount of removal
Box 5.1} (medium confidence). {4.6.3, 5.6.2.1, Figure 5.32, Figure 5.34}
TS.3.3.2 Carbon Dioxide Removal Due to non-linearities in the climate system, the century-scale
climate–carbon cycle response to a CO2 removal from the atmosphere
Deliberate carbon dioxide removal (CDR) from the is not always equal and opposite to its response to a simultaneous
atmosphere has the potential to compensate for residual CO2 emission (medium confidence). For CO2 emissions of 100 PgC
CO2 emissions to reach net zero CO2 emissions or to released from a state in equilibrium with pre-industrial atmospheric
generate net negative CO2 emissions. In the same way CO2 levels, CMIP6 models simulate that 27± 6% (mean ± 1 standard
that part of current anthropogenic net CO2 emissions deviation) of emissions remain in the atmosphere 80–100 years
are taken up by land and ocean carbon stores, net CO2 after the emissions, whereas for removals of 100 PgC only 23
removal will be partially counteracted by CO2 release from ± 6% of removals remain out of the atmosphere. This asymmetry
these stores (very high confidence). Asymmetry in the implies that an extra amount of CDR is required to compensate for
carbon cycle response to simultaneous CO2 emissions and a positive emission of a given magnitude to attain the same change
removals implies that a larger amount of CO2 would need in atmospheric CO2. Due to low agreement between models, there
99
Technical Summary
is low confidence in the sign of the asymmetry of the temperature Carbon dioxide removal methods have a range of side effects that
response to CO2 emissions and removals. {4.6.3, 5.6.2.1, Figure 5.35} can either weaken or strengthen the carbon sequestration and
cooling potential of these methods and affect the achievement
Simulations with ESMs indicate that under scenarios where CO2 of sustainable development goals (high confidence). Biophysical
emissions gradually decline, reach net zero and become net negative and biogeochemical side-effects of CDR methods are associated
during the 21st century (e.g., SSP1-2.6), land and ocean carbon sinks with changes in surface albedo, the water cycle, emissions of CH4
begin to weaken in response to declining atmospheric CO2 and N2O, ocean acidification and marine ecosystem productivity
concentrations, and the land sink eventually turns into a source (high confidence). These side-effects and associated Earth system
(Figure TS.19). This sink-to-source transition occurs decades to a few feedbacks can decrease carbon uptake and/or change local and
centuries after CO2 emissions become net negative. The ocean regional climate and in turn limit the CO2 sequestration and cooling
remains a sink of CO2 for centuries after emissions become net potential of specific CDR methods (medium confidence). Deployment
negative. Under scenarios with large net negative CO2 emissions of CDR, particularly on land, can also affect water quality and quantity,
(e.g., SSP5-3.4-OS) and rapidly declining CO2 concentrations, the land food production and biodiversity (high confidence). These effects are
source is larger than for SSP1-2.6 and the ocean also switches to a often highly dependent on local context, management regime, prior
source. While the general response is robust across models, there is land use, and scale (high confidence). The largest co-benefits are
low confidence in the timing of the sink-to-source transition and the obtained with methods that seek to restore natural ecosystems or
magnitude of the CO2 source in scenarios with net negative CO2 improve soil carbon sequestration (medium confidence). The climate
emissions. Carbon dioxide removal could reverse some aspects and biogeochemical effects of terminating CDR are expected to be
climate change if CO2 emissions become net negative, but some small for most CDR methods (medium confidence). {4.6.3, 5.6.2.2,
changes would continue in their current direction for decades to Figure 5.36, 8.4.3, 8.6.3}
millennia. For instance, sea level rise due to ocean thermal expansion
would not reverse for several centuries to millennia (high confidence)
TS (Box TS.4). {4.6.3, 5.4.10, 5.6.2.1, Figure 5.30, Figure 5.33}
550
Net ocean flux
Carbon dioxide concentration (ppm)
203 –41
500
–61 33 –23
–33
and removals
450
Net CO2 emissions
–30 2
4 3 0
400 469
446 –7 –5 –7
411 –7
403 400 396
350 368
+101 ppm –23 ppm –35 ppm –8 ppm –4 ppm
300
2000 2050 2100 2150 2200 2250 2300
Figure TS.19 | Carbon sink response in a scenario with net carbon dioxide (CO2) removal from the atmosphere. The intent of this figure is to show how
atmospheric CO2 evolves under negative emissions and its dependence on the negative emissions technologies. It also shows the evolution of the ocean and land sinks. Shown
are CO2 flux components from concentration-driven Earth system model (ESM) simulations during different emissions stages of SSP1–2.6 and its long-term extension. (a) Large
net positive CO2 emissions, (b) small net positive CO2 emissions, (c–d) net negative CO2 emissions, and (e) net zero CO2 emissions. Positive flux components act to raise the
atmospheric CO2 concentration, whereas negative components act to lower the CO2 concentration. Net CO2 emissions and land and ocean CO2 fluxes represent the multi-model
mean and standard deviation (error bar) of four ESMs (CanESM5, UKESM1, CESM2-WACCM, IPSL-CM6a-LR) and one Earth system model of intermediate complexity (Uvic
ESCM). Net CO2 emissions are calculated from concentration-driven ESM simulations as the residual from the rate of increase in atmospheric CO2 and land and ocean CO2 fluxes.
Fluxes are accumulated over each 50-year period and converted to concentration units (parts per million, or ppm). {5.6.2.1, Figure 5.33}
100
Technical Summary
TS.3.3.3 Relating Different Forcing Agents future global surface temperature outcomes (high confidence) {7.6.1,
Box 7.3}
When including other GHGs, the choice of emissions metric
affects the quantification of net zero GHG emissions and Emissions metrics are needed to aggregate baskets of gases to
their resulting temperature outcome (high confidence). determine net zero GHG emissions. Generally, achieving net zero CO2
Reaching and sustaining net zero GHG emissions typically emissions and declining non-CO2 radiative forcing would halt human-
leads to a peak and decline in temperatures when induced warming. Reaching net zero GHG emissions quantified by
quantified with the global warming potential over a 100- GWP-100 typically leads to declining temperatures after net zero
year period (GWP-100). Carbon-cycle responses are more GHGs emissions are achieved if the basket includes short-lived gases,
robustly accounted for in emissions metrics compared to such as CH4. Net zero GHG emissions defined by CGTP or GWP*
AR5 (high confidence). New emissions metric approaches imply net zero CO2 and other long-lived GHG emissions and constant
can be used to generate equivalent cumulative emissions (CGTP) or gradually declining (GWP*) emissions of short-lived gases.
of CO2 for short-lived greenhouse gases based on their rate The warming evolution resulting from net zero GHG emissions
of emissions. {7.6.2} defined in this way corresponds approximately to reaching net zero
CO2 emissions, and would thus not lead to declining temperatures
Over 10- to 20-year time scales, the temperature response after net zero GHG emissions are achieved but to an approximate
to a single year’s worth of current emissions of short- temperature stabilization (high confidence). The choice of emissions
lived climate forcers (SLCFs) is at least as large as that of metric hence affects the quantification of net zero GHG emissions,
CO2, but because the effects of SLCFs decay rapidly over and therefore the resulting temperature outcome of reaching and
the first few decades after emission, the net long-term sustaining net zero GHG emissions levels (high confidence). {7.6.1.4,
temperature response to a single year’s worth of emissions 7.6.2, 7.6.3}
is predominantly determined by cumulative CO2 emissions.
As pointed out in AR5, ultimately, it is a matter for policymakers TS
Emissions reductions in 2020 associated with COVID-19 to decide which emissions metric is most applicable to their needs.
containment led to small and positive global ERF; however, This Report does not recommend the use of any specific emissions
global and regional climate responses to the forcing are metric, as the most appropriate metric depends on the policy goal
undetectable above internal variability due to the temporary and context (see Chapter 7, Section 7.6). A detailed assessment of
nature of emissions reductions. {6.6, Cross-Chapter Box 6.1} GHG metrics to support climate change mitigation and associated
policy contexts is provided in the WGIII contribution to the AR6.
The relative climate effects of different forcing agents are typically
quantified using emissions metrics that compare the effects of The global surface temperature response following a climate change
an idealised pulse of 1 kg of some climate forcing agent against mitigation measure that affects emissions of both short- and long-
a reference climate forcing agent, almost always CO2. The two most lived climate forcers depends on their lifetimes, their ERFs, how
prominent pulse emissions metrics are the global warming potential fast and for how long the emissions are reduced, and the thermal
(GWP) and global temperature change potential (GTP) (see Glossary). inertia in the climate system. Mitigation, relying on emissions
The climate responses to CO2 emissions by convention include the reductions and implemented through new legislation or technology
effects of warming on the carbon cycle, so for consistency these also standards, implies that emissions reductions occur year after year.
need to be determined for non-CO2 emissions. The methodology Global temperature response to a year’s worth of current emissions
for doing this has been placed on a more robust scientific footing from different sectors informs about the mitigation potential (Figure
compared to AR5 (high confidence). Methane from fossil fuel sources TS.20). Over 10- to 20-year time scales, the influence of SLCFs is
has slightly higher emissions metric values than those from biogenic at least as large as that of CO2, with sectors producing the largest
sources since it leads to additional fossil CO2 in the atmosphere warming being fossil fuel production and distribution, agriculture,
(high confidence). Updates to the chemical adjustments for CH4 and and waste management. Because the effects of the SLCFs decay
N2O emissions (Section TS.3.1) and revisions in their lifetimes result rapidly over the first few decades after emission, the net long-term
in emissions metrics for GWP and GTP that are slightly lower than temperature effect from a single year’s worth of current emissions is
in AR5 (medium confidence). Emissions metrics for the entire suite predominantly determined by CO2. Fossil fuel combustion for energy,
of GHGs assessed in the AR6 have been calculated for various time industry and land transportation are the largest contributing sectors
horizons. {7.6.1, Table 7.15, Table 7.SM.7} on a 100-year time scale (high confidence). Current emissions of CO2,
N2O and SLCFs from East Asia and North America are the largest
New emissions metric approaches, such as GWP* and Combined- regional contributors to additional net future warming on both
GTP (CGTP), relate changes in the emissions rate of short-lived short (medium confidence) and long time scales (10 and 100 years,
greenhouse gases to equivalent cumulative emissions of CO2 respectively) (high confidence). {6.6.1, 6.6.2, Figure 6.16}
(CO2-e). Global surface temperature response from aggregated
emissions of short-lived greenhouse gases over time is determined COVID-19 restrictions led to detectable reductions in global
by multiplying these cumulative CO2-e by TCRE (see Section TS.3.2.1). anthropogenic emissions of nitrogen oxides (NOx) (about 35% in
When GHGs are aggregated using standard metrics such as GWP or April 2020) and fossil CO2 (7%, with estimates ranging from 5.8% to
GTP, cumulative CO2-e emissions are not necessarily proportional to 13.0%), driven largely by reduced emissions from the transportation
101
Technical Summary
−0.06 −0.05 −0.04 −0.03 −0.02 −0.01 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Change in global surface temperature due to total anthropogenic emissions (°C)
−0.010 −0.005 0 0.005 0.010 0.015 −0.010 −0.005 0 0.005 0.010 0.015
Change in global surface temperature (°C) Change in global surface temperature (°C)
Figure TS.20 | Global surface temperature change 10 and 100 years after a one-year pulse of present-day emissions. The intent of this figure is to show the
sectoral contribution to present-day climate change by specific climate forcers, including carbon dioxide (CO2) as well as short-lived climate forcers (SLCFs). The temperature
response is broken down by individual species and shown for total anthropogenic emissions (top), and sectoral emissions on 10-year (left) and 100-year time scales (right).
Sectors are sorted by (high-to-low) net temperature effect on the 10-year time scale. Error bars in the top panel show the 5–95% range in net temperature effect due to
uncertainty in radiative forcing only (calculated using a Monte Carlo approach and best estimate uncertainties from the literature). Emissions for 2014 are from the Coupled
Model Intercomparison Project Phase 6 (CMIP6) emissions dataset, except for hydrofluorocarbons (HFCs) and aviation H2O, which rely on other datasets (see Section 6.6.2 for
more details). CO2 emissions are excluded from open biomass burning and residential biofuel use. {6.6.2, Figure 6.16}
sector (medium confidence). There is high confidence that, with temporarily adding to the total anthropogenic climate influence,
the exception of surface ozone, reductions in pollutant precursors with positive forcing (warming influence) from aerosol changes
contributed to temporarily improved air quality in most regions of dominating over negative forcings (cooling influence) from CO2, NOx
the world. However, these reductions were lower than what would and contrail cirrus changes. Consistent with this small net radiative
be expected from sustained implementation of policies addressing forcing, and against a large component of internal variability, Earth
air quality and climate change (medium confidence). Overall, the net system models show no detectable effect on global or regional
global ERF from COVID-19 containment was likely small and positive surface temperature or precipitation (high confidence). {Cross
for 2020 (with a temporary peak value less than 0.2 W m–2), thus Chapter Box 6.1}
102
Technical Summary
Box TS.7 | Climate and Air Quality Responses to Short-lived Climate Forcers in Shared
Socio-economic Pathways
Future changes in emissions of short-lived climate forcers (SLCFs) are expected to cause an additional global mean
warming, with a large diversity in the end-of-century response across the WGI core set of Shared Socio-economic
Pathways (SSPs), depending upon the level of climate change and air pollution mitigation (Box TS.7, Figure 1). This
additional warming is either due to reductions in cooling aerosols for air pollution regulation or due to increases in
methane (CH4), ozone and hydrofluorocarbons (HFCs). This additional warming is stable after 2040 in SSPs associated
with lower global air pollution as long as CH4 emissions are also mitigated, but the overall warming induced by SLCF
changes is higher in scenarios in which air quality continues to deteriorate (induced by growing fossil fuel use and
limited air pollution control) (high confidence).
Sustained CH4 mitigation reduces global surface ozone, contributing to air quality improvements, and also reduces
surface temperature in the longer term, but only sustained CO2 emissions reductions allow long-term climate
stabilization (high confidence). Future changes in air quality (near-surface ozone and particulate matter, or PM) at
global and local scales are predominantly driven by changes in ozone and aerosol precursor emissions rather than
climate (high confidence). Air quality improvements driven by rapid decarbonization strategies, as in SSP1-1.9 and
SSP1-2.6, are not sufficient in the near term to achieve air quality guidelines set by the World Health Organization in
some highly polluted regions (high confidence). Additional policies (e.g., access to clean energy, waste management)
envisaged to attain United Nations Sustainable Development Goals bring complementary SLCF reduction. {4.4.4, 6.6.3,
6.7.3, Box 6.2}
TS
The net effect of SLCF emissions changes on temperature will depend on how emissions of warming and cooling SLCFs will evolve
in the future. The magnitude of the cooling effect of aerosols remains the largest uncertainty in the effect of SLCFs in future climate
projections. Since the SLCFs have undergone large changes over the past two decades, the temperature and air pollution responses
are estimated relative to the year 2019 instead of 1995–2014.
Temperature Response
In the next two decades, it is very likely that SLCF emissions changes will cause a warming relative to 2019, across the WGI core set of
SSPs (see Section TS.1.3.1), in addition to the warming from long-lived GHGs. The net effect of SLCF and HFC changes in global surface
temperature across the SSPs is a likely warming of 0.06°C–0.35°C in 2040 relative to 2019. This near-term global mean warming
linked to SLCFs is quite similar in magnitude across the SSPs due to competing effects of warming (CH4, ozone) and cooling (aerosols)
forcers (Box TS.7, Figure 1). There is greater diversity in the end-of-century response among the scenarios. SLCF changes in scenarios
with no climate change mitigation (SSP3-7.0 and SSP5-8.5) will cause a warming in the likely range of 0.4°C–0.9°C in 2100 relative to
2019 due to increases in CH4, tropospheric ozone and HFC levels. For the stringent climate change and pollution mitigation scenarios
(SSP1-1.9 and SSP1-2.6), the cooling from reductions in CH4, ozone and HFCs partially balances the warming from reduced aerosols,
primarily sulphate, and the overall SLCF effect is a likely increase in global surface temperature of 0.0°C–0.3°C in 2100, relative to
2019. With intermediate climate change and air pollution mitigations, SLCFs in SSP2-4.5 add a likely warming of 0.2°C–0.5°C to global
surface temperature change in 2100, with the largest warming resulting from reductions in aerosols. {4.4.4, 6.7.3}
Assuming implementation and efficient enforcement of both the Kigali Amendment to the Montreal Protocol on Substances that
Deplete the Ozone Layer and current national plans result in limiting emissions (as in SSP1-2.6), the effects of HFCs on global
surface temperature, relative to 2019, would remain below +0. 02°C from 2050 onwards versus about +0.04°C–0.08°C in 2050 and
+0.1°C–0.3°C in 2100 considering only national HFC regulations decided prior to the Kigali Amendment (as in SSP5-8.5) (medium
confidence). {6.6.3, 6.7.3}
103
Technical Summary
TS
Box TS.7, Figure 1 | Effects of short-lived climate forcers (SLCFs) on global surface temperature and air pollution across the WGI core set of
Shared Socio-economic Pathways (SSPs). The intent of this figure is to show the climate and air quality (surface ozone and particulate matter smaller than
2.5 microns in diameter, or PM2.5 ) response to SLCFs in the SSP scenarios for the near and long-term. Effects of net aerosols, tropospheric ozone, hydrofluorocarbons
(HFCs; with lifetimes less than 50 years), and methane (CH4) are compared with those of total anthropogenic forcing for 2040 and 2100 relative to year 2019. The
global surface temperature changes are based on historical and future evolution of effective radiative forcing (ERF) as assessed in Chapter 7 of this Report. The
temperature responses to the ERFs are calculated with a common impulse response function (RT) for the climate response, consistent with the metric calculations in
Chapter 7 (Box 7.1). The RT has an equilibrium climate sensitivity of 3.0°C for a doubling of atmospheric CO2 concentration (feedback parameter of –1.31 W m–2 °C–1).
The scenario total (grey bar) includes all anthropogenic forcings (long- and short-lived climate forcers, and land-use changes). Uncertainties are 5–95% ranges. The
global changes in air pollutant concentrations (ozone and PM2.5) are based on multimodel Coupled Model Intercomparison Project Phase 6 (CMIP6) simulations and
represent changes in five-year mean surface continental concentrations for 2040 and 2098 relative to 2019. Uncertainty bars represent inter-model ±1 standard
deviation. {6.7.2, 6.7.3, Figure 6.24}
21 Although cirrus cloud thinning aims to cool the planet by increasing longwave emissions to space, it is included in the portfolio of SRM options for consistency with AR5 and SR1.5. {4.6.3.3}
104
Technical Summary
Solar radiation modification (SRM) refers to deliberate, large-scale climate intervention options that are studied as potential
supplements to deep mitigation, for example, in scenarios that overshoot climate stabilization goals. SRM options aim to offset some
of the warming effects of GHG emissions by modification of Earth’s shortwave radiation budget. Following SR1.5, the SRM assessed
in this Report also includes some options, such as cirrus cloud thinning, that alter the longwave radiation budget.
SRM contrasts with climate change mitigation activities, such as emissions reductions and CDR, as it introduces a ‘mask’ to the
climate change problem by altering Earth’s radiation budget, rather than attempting to address the root cause of the problem, which
is the increase in GHGs in the atmosphere. By masking only the climate effects of GHG emissions, SRM does not address other issues
related to atmospheric CO2 increase, such as ocean acidification. This Report assesses physical understanding of the Earth system
response to proposed SRM, and the assessment is based primarily on idealized climate model simulations. There are other important
considerations, such as risk to human and natural systems, perceptions, ethics, cost, governance, and trans-boundary issues and their
relationship to the United Nations Sustainable Development Goals – issues that the WGII (Chapter 16) and WGIII (Chapter 14) Reports
address. {4.6.3}
SRM options include those that increase surface albedo, brighten marine clouds by increasing the amount of cloud condensation
nuclei, or reduce the optical depth of cirrus clouds by seeding them with ice nucleating particles. However, the most commonly studied
approaches attempt to mimic the cooling effects of major volcanic eruptions by injecting reflective aerosols (e.g., sulphate aerosols)
or their precursors (e.g., sulphur dioxide) into the stratosphere. {4.6.3, 5.6.3, 6.4.6}
SRM could offset some effects of greenhouse gas-induced warming on global and regional climate, but there would be substantial TS
residual and overcompensating climate change at the regional scale and seasonal time scales (high confidence). Since AR5, more
modelling work has been conducted with more sophisticated treatment of aerosol-based SRM approaches, but the uncertainties in
cloud–aerosol–radiation interactions are still large (high confidence). Modelling studies suggest that it is possible to stabilize multiple
large-scale temperature indicators simultaneously by tailoring the deployment strategy of SRM options (medium confidence) but with
large residual or overcompensating regional and seasonal climate changes. {4.6.3}
SRM approaches targeting shortwave radiation are likely to reduce global mean precipitation, relative to future CO2 emissions
scenarios, if all global mean warming is offset. In contrast, cirrus cloud thinning, targeting longwave radiation, is expected to cause
an increase in global mean precipitation (medium confidence). If shortwave approaches are used to offset global mean warming,
the magnitude of reduction in regional precipitation minus evapotranspiration (P–E) (Box TS.5), which is more relevant to freshwater
availability, is smaller than precipitation decrease because of simultaneous reductions in both precipitation and evapotranspiration
(medium confidence). {4.6.3, 8.2.1, 8.6.3}.
If SRM is used to cool the planet, it would cause a reduction in plant and soil respiration and slow the reduction of ocean carbon
uptake due to warming (medium confidence). The result would be an enhancement of the global land and ocean CO2 sinks (medium
confidence) and a slight reduction in atmospheric CO2 concentration relative to unmitigated climate change. However, SRM would
not stop CO2 from increasing in the atmosphere or affect the resulting ocean acidification under continued anthropogenic emissions
(high confidence). {5.6.3}
The effect of stratospheric aerosol injection on global temperature and precipitation is projected by models to be detectable after one
to two decades, which is similar to the time scale for the emergence of the benefits of emissions reductions. A sudden and sustained
termination of SRM in a high GHG emissions scenario would cause rapid climate change and a reversal of the SRM effects on the
carbon sinks (high confidence). It is also likely that a termination of strong SRM would drive abrupt changes in the water cycle globally
and regionally, especially in the tropical regions by shifting the Inter-tropical Convergence Zone and Hadley cells. At the regional scale,
non-linear responses cannot be excluded, due to changes in evapotranspiration. However, a gradual phase-out of SRM combined with
emissions reductions and CDR would avoid larger rates of changes (medium confidence). {4.6.3, 5.6.3, 8.6.3}.
105
Technical Summary
The present rates of response of many aspects of the climate system are proportionate to the rate of recent temperature
change, but some aspects may respond disproportionately. Some climate system components are slow to respond,
such as the deep ocean overturning circulation and the ice sheets (Box TS.4). It is virtually certain that irreversible,
committed change is already underway for the slow-to-respond processes as they come into adjustment for past and
present emissions.
The paleoclimate record indicates that tipping elements exist in the climate system where processes undergo sudden
shifts toward a different sensitivity to forcing, such as during a major deglaciation, where 1°C degree of temperature
change might correspond to a large or small ice-sheet mass loss during different stages (Box TS.2). For global climate
indicators, evidence for abrupt change is limited, but deep ocean warming, acidification and sea level rise are committed
to ongoing change for millennia after global surface temperatures initially stabilize and are irreversible on human time
scales (very high confidence). At the regional scale, abrupt responses, tipping points and even reversals in the direction of
change cannot be excluded (high confidence). Some regional abrupt changes and tipping points could have severe local
impacts, such as unprecedented weather, extreme temperatures and increased frequency of droughts and forest fires.
Models that exhibit such tipping points are characterized by abrupt changes once the threshold is crossed, and even
a return to pre-threshold surface temperatures or to atmospheric carbon dioxide concentrations does not guarantee
that the tipping elements return to their pre-threshold state. Monitoring and early warning systems are being put into
place to observe tipping elements in the climate system. {1.3, 1.4.4, 1.5, 4.3.2, Table 4.10, 5.3.4, 5.4.9, 7.5.3, 9.2.2,
TS 9.2.4, 9.4.1, 9.4.2, 9.6.3, Cross-chapter Box 12.1}
Understanding of multi-decadal reversibility (i.e., the system returns to the previous climate state within multiple decades after
the radiative forcing is removed) has improved since AR5 for many atmospheric, land surface and sea ice climate metrics following
sea surface temperature recovery. Some processes suspected of having tipping points, such as the Atlantic Meridional Overturning
Circulation (AMOC), have been found to often undergo recovery after temperature stabilization with a time delay (low confidence).
However, substantial irreversibility is further substantiated for some cryosphere changes, ocean warming, sea level rise, and ocean
acidification. {4.7.2, 5.3.3, 5.4.9, 9.2.2, 9.2.4, 9.4.1, 9.4.2, 9.6.3}
Some climate system components are slow to respond, such as the deep ocean overturning circulation and the ice sheets. It is likely that
under stabilization of global warming at 1.5°C, 2.0°C or 3.0°C relative to 1850–1900, the AMOC will continue to weaken for several
decades by about 15%, 20% and 30% of its strength and then recover to pre-decline values over several centuries (medium confidence).
At sustained warming levels between 2°C and 3°C, there is limited evidence that the Greenland and West Antarctic ice sheets will be lost
almost completely and irreversibly over multiple millennia; both the probability of their complete loss and the rate of mass loss increases
with higher surface temperatures (high confidence). At sustained warming levels between 3°C and 5°C, near-complete loss of the
Greenland Ice Sheet and complete loss of the West Antarctic Ice Sheet is projected to occur irreversibly over multiple millennia (medium
confidence); with substantial parts or all of Wilkes Subglacial Basin in East Antarctica lost over multiple millennia (low confidence). Early-
warning signals of accelerated sea level rise from Antarctica could possibly be observed within the next few decades. For other hazards
(e.g., ice-sheet behaviour, glacier mass loss and global mean sea level change, coastal floods, coastal erosion, air pollution, and ocean
acidification) the time and/or scenario dimensions remain critical, and a simple and robust relationship with global warming level cannot
be established (high confidence). {4.3.2, 4.7.2, 5.4.3, 5.4.5, 5.4.8, 8.6, 9.2, 9.4, Box 9.3, Cross-Chapter Box 12.1}
For global climate indicators, evidence for abrupt change is limited. For global warming up to 2°C above 1850–1900 levels, paleoclimate
records do not indicate abrupt changes in the carbon cycle (low confidence). Despite the wide range of model responses, uncertainty
in atmospheric CO2 by 2100 is dominated by future anthropogenic emissions rather than uncertainties related to carbon–climate
feedbacks (high confidence). There is no evidence of abrupt change in climate projections of global temperature for the next century:
there is a near-linear relationship between cumulative CO2 emissions and maximum global mean surface air temperature increase
caused by CO2 over the course of this century for global warming levels up to at least 2°C relative to 1850–1900. The increase in global
ocean heat content (Section TS.2.4) will likely continue until at least 2300 even for low emissions scenarios, and global mean sea level
will continue to rise for centuries to millennia following cessation of emissions (Box TS.4) due to continuing deep ocean heat uptake
and mass loss of the Greenland and Antarctic ice sheets (high confidence). {2.2.3; Cross-Chapter Box 2.1; 5.1.1; 5.4; Cross-Chapter
Box 5.1; Figures 5.3, 5.4, 5.25, and 5.26; 9.2.2; 9.2.4}
The response of biogeochemical cycles to anthropogenic perturbations can be abrupt at regional scales and irreversible on decadal to
century time scales (high confidence). The probability of crossing uncertain regional thresholds increases with climate change (high
106
Technical Summary
confidence). It is very unlikely that gas clathrates (mostly methane) in deeper terrestrial permafrost and subsea clathrates will lead to a
detectable departure from the emissions trajectory during this century. Possible abrupt changes and tipping points in biogeochemical
cycles lead to additional uncertainty in 21st century atmospheric GHG concentrations, but future anthropogenic emissions remain the
dominant uncertainty (high confidence). There is potential for abrupt water cycle changes in some high emissions scenarios, but there
is no overall consistency regarding the magnitude and timing of such changes. Positive land surface feedbacks, including vegetation,
dust, and snow, can contribute to abrupt changes in aridity, but there is only low confidence that such changes will occur during
the 21st century. Continued Amazon deforestation, combined with a warming climate, raises the probability that this ecosystem will
cross a tipping point into a dry state during the 21st century (low confidence). (Section TS.3.2.2) {5.4.3, 5.4.5, 5.4.8, 5.4.9, 8.6.2, 8.6.3,
Cross-Chapter Box 12.1}
TS.4 Regional Climate Change requirements, and demand (very high confidence). The
decision-making context, level of user engagement,
This section focuses on how to generate regional climate change and co-production between scientists, practitioners and
information and its relevance for climate services; the drivers of regional users are important determinants of the type of climate
climate variability and change and how they are being affected by service developed and its utility in supporting adaptation,
anthropogenic factors; and observed, attributed and projected changes mitigation and risk management decisions. {10.3, 10.6,
in climate, including extreme events and climatic impact-drivers (CIDs), Cross-Chapter Box 10.3, 12.6, Cross-Chapter Box 12.2}
across all regions of the world. There is a small set of CID changes TS
common to all land or ocean regions and a specific set of changes from
a broader range of CIDs seen in each region. This regional diversity TS.4.1.1 Sources and Methodologies for Generating Regional
results from regional climate being determined by a complex interplay Climate Information
between the seasonal-to-multi-decadal variation of large-scale modes
of climate variability, external natural and anthropogenic forcings, local Climate change information at regional scale is generated using a range
climate processes and related feedbacks. of data sources and methodologies (Section TS.1.4). Understanding
of observed regional climate change and variability is based on the
availability and analysis of multiple observational datasets that are
TS.4.1 Generation and Communication of Regional suitable for evaluating the phenomena of interest (e.g., extreme
Climate Change Information events), including accounting for observational uncertainty (Section
TS.1.2.1). These datasets are combined with climate model simulations
Climate change information at regional scale is generated of observed changes and events to attribute causes of those changes
using a range of data sources and methodologies. Multi- and events to large- and regional-scale anthropogenic and natural
model ensembles and models with a range of resolutions drivers and to assess the performance of the models. Future simulations
are important data sources, and discarding models that with many climate models (multi-model ensembles) are then used to
fundamentally misrepresent relevant processes improves generate and quantify ranges of projected regional climate responses
the credibility of ensemble information related to these (Section TS.4.2). Discarding models that fundamentally misrepresent
processes. A key methodology is distillation – combining relevant processes improves the credibility of regional climate
lines of evidence and accounting for stakeholder context information generated from these ensembles (high confidence).
and values – which helps ensure the information is relevant, However, multi-model mean and ensemble spread are not a full
useful and trusted for decision-making (see Core Concepts measure of the range of projection uncertainty and are not sufficient
Box) (high confidence). to characterize low-likelihood, high-impact changes (Box TS.3) or
situations where different models simulate substantially different
Since AR5, physical climate storylines have emerged as or even opposite changes (high confidence). Large single-model
a complementary approach to ensemble projections for ensembles are now available and provide a more comprehensive
generating more accessible climate information and spectrum of possible changes associated with internal variability (high
promoting a more comprehensive treatment of risk. They confidence) (Section TS.1.2.3). {1.5.1, 1.5.4, 10.2, 10.3.3, 10.3.4, 10.4.1,
have been used as part of the distillation process within 10.6.2, 11.2, Box 11.2, Cross-Chapter Box 11.1, 12.4, Atlas.1.4.1}
climate services to generate the required context-relevant,
credible and trusted climate information. Depending on the region of interest, representing regionally important
forcings (e.g., aerosols, land-use change and ozone concentrations)
Since AR5, climate change information produced for climate and feedbacks (e.g., between snow and albedo, soil moisture and
services has increased significantly due to scientific and temperature, or soil moisture and precipitation) in climate models is a
technological advancements and growing user awareness, prerequisite for them to reproduce past regional trends to underpin the
107
Technical Summary
reliability of future projections (medium confidence) (Section TS.1.2.2). Methodologies such as statistical downscaling, bias adjustment and
In some cases, even the sign of a projected change in regional climate weather generators are beneficial as an interface between climate
cannot be trusted if relevant regional processes are not represented, for model projections and impact modelling and for deriving user-
example, for variables such as precipitation and wind speed (medium relevant indicators (high confidence). However, the performance of
confidence). In some regions, either geographical (e.g., Central Africa, these techniques depends on that of the driving climate model: in
Antarctica) or typological (e.g., mountainous areas, Small Islands and particular, bias adjustment cannot overcome all consequences of
cities), and for certain phenomena, fewer observational records are unresolved or strongly misrepresented physical processes, such as
available or accessible, which limits the assessment of regional climate large-scale circulation biases or local feedbacks (medium confidence).
change in these cases. {1.5.1, 1.5.3, 1.5.4, 8.5.1, 10.2, 10.3.3, 10.4.1, {10.3.3, Cross-Chapter Box 10.2, 12.2, Atlas.2.2}
11.1.6, 11.2, 12.4, Atlas.8.3, Atlas.11.1.5, Cross-Chapter Box Atlas.2}
The attribution of observed changes in extremes to human influence (including greenhouse gas and aerosol emissions
and land-use changes) has substantially advanced since AR5, in particular for extreme precipitation, droughts,
tropical cyclones, and compound extremes (high confidence). There is limited evidence for windstorms and convective
storms. Some recent hot extreme events would have been extremely unlikely to occur without human influence on the
climate system. (Section TS.1) {Cross-Working Group Box: Attribution in Chapter 1, 11.2, 11.3, 11.4, 11.6, 11.7, 11.8}
Since AR5, the attribution of extreme weather events has emerged as a growing field of climate research with an increasing body of
TS literature. It provides evidence that greenhouse gases and other external forcings have affected individual extreme weather events
by disentangling anthropogenic drivers from natural variability. Event attribution is now an important line of evidence for assessing
changes in extremes on regional scales. (Section TS.1) {Cross-Working Group Box: Attribution, 11.1.4}
The regional extremes and events that have been studied are geographically uneven (Section TS.4.1). A few events, for example,
extreme rainfall events in the United Kingdom, heatwaves in Australia, or Hurricane Harvey that hit Texas in 2017, have been heavily
studied. Many highly impactful extreme weather events have not been studied in the event attribution framework, particularly in
the developing world where studies are generally lacking. This is due to various reasons, including lack of observational data, lack of
reliable climate models, and lack of scientific capacity. While the events that have been studied are not representative of all extreme
events that have occurred, and results from these studies may also be subject to selection bias, the large number of event attribution
studies provide evidence that changes in the properties of these local and individual events are in line with expected consequences
of human influence on the climate and can be attributed to external drivers. {Cross-Working Group Box: Attribution, 11.1.4, 11.2.2}
It is very likely that human influence is the main contributor to the observed increase in the intensity and frequency of hot extremes and
the observed decrease in the intensity and frequency of cold extremes on continental scales. Some specific recent hot extreme events
would have been extremely unlikely to occur without human influence on the climate system. Changes in aerosol concentrations
have likely slowed the increase in hot extremes in some regions, in particular from 1950–1980. No-till farming, irrigation and crop
expansion have similarly attenuated increases in summer hot extremes in some regions, such as central North America (medium
confidence). {11.3.4}
Human influence has contributed to the intensification of heavy precipitation in three continents where observational data are most
abundant: North America, Europe and Asia (high confidence). On regional scales, evidence of human influence on extreme precipitation
is limited, but new evidence from attributing individual heavy precipitation events found that human influence was a significant driver
of the events. {11.4.4}
There is low confidence that human influence has affected trends in meteorological droughts in most regions, but medium confidence
that they have contributed to the severity of some specific events. There is medium confidence that human-induced climate change
has contributed to increasing trends in the probability or intensity of recent agricultural and ecological droughts, leading to an
increase of the affected land area. {11.6.4}
Event attribution studies of specific strong tropical cyclones provide limited evidence for anthropogenic effects on tropical cyclone
intensifications so far, but high confidence for increases in precipitation. There is high confidence that anthropogenic climate change
contributed to extreme rainfall amounts during Hurricane Harvey (in 2017) and other intense tropical cyclones. {11.7.3}
108
Technical Summary
The number of evident attribution studies on compound events is limited. There is medium confidence that weather conditions that
promote wildfires have become more probable in southern Europe, northern Eurasia, the USA, and Australia over the last century. In
Australia a number of event attribution studies show that there is medium confidence of increase in fire weather conditions due to
human influence. {11.8.3, 12.4.3.2}
Climate change is already affecting every inhabited region across the globe,
with human influence contributing to many observed changes in weather
and climate extremes
(a) Synthesis of assessment of observed change in hot extremes and
confidence in human contribution to the observed changes in the world’s regions
Type of observed change
in hot extremes North
America GIC Europe
Increase (41) NWN NEN NEU RAR
Each hexagon corresponds IPCC AR6 WGI reference regions: North America: NWN (North-Western North America, NEN (North-Eastern North America), WNA
to one of the IPCC AR6 (Western North America), CNA (Central North America), ENA (Eastern North America), Central America: NCA (Northern Central America),
WGI reference regions SCA (Southern Central America), CAR (Caribbean), South America: NWS (North-Western South America), NSA (Northern South America), NES
(North-Eastern South America), SAM (South American Monsoon), SWS (South-Western South America), SES (South-Eastern South America),
North-Western SSA (Southern South America), Europe: GIC (Greenland/Iceland), NEU (Northern Europe), WCE (Western and Central Europe), EEU (Eastern
NWN Europe), MED (Mediterranean), Africa: MED (Mediterranean), SAH (Sahara), WAF (Western Africa), CAF (Central Africa), NEAF (North Eastern
North America
Africa), SEAF (South Eastern Africa), WSAF (West Southern Africa), ESAF (East Southern Africa), MDG (Madagascar), Asia: RAR (Russian
Arctic), WSB (West Siberia), ESB (East Siberia), RFE (Russian Far East), WCA (West Central Asia), ECA (East Central Asia), TIB (Tibetan Plateau),
EAS (East Asia), ARP (Arabian Peninsula), SAS (South Asia), SEA (South East Asia), Australasia: NAU (Northern Australia), CAU (Central
Australia), EAU (Eastern Australia), SAU (Southern Australia), NZ (New Zealand), Small Islands: CAR (Caribbean), PAC (Pacific Small Islands)
Box TS.10, Figure 1 | Synthesis of assessed observed and attributable regional changes.
109
Technical Summary
TS
TS.4.1.2 Regional Climate Information Distillation and changes in a key mode of variability (the Southern Annular Mode)
Climate Services and drought in Cape Town among different observation periods and
in model simulations. {10.5.3, 10.6, 10.6.2, 10.6.4, Cross-Chapter
The construction of regional climate information involves people with Box 10.3, 12.4}
a variety of backgrounds, from various disciplines, who have different
sets of experiences, capabilities and values. The process of synthesizing Since AR5, physical climate storyline approaches have emerged as
climate information from different lines of evidence from a number a complementary instrument to provide a different perspective, or
of sources, taking into account the context of a user vulnerable to additional climate information, to facilitate communication of the
climate variability and change and the values of all relevant actors, is information or provide a more flexible consideration of risk. Storylines
called distillation. Distillation is conditioned by the sources available, that condition climatic events and processes on a set of plausible
the actors involved, and the context, which all depend heavily on the but distinct large-scale climatic changes enable the exploration of
regions considered, and is framed by the question being addressed. uncertainties in regional climate projections. For example, they can
Distilling regional climate information from multiple lines of evidence explicitly address low-likelihood, high-impact outcomes, which would
and taking the user context into account increases fitness, usefulness, be less emphasized in a probabilistic approach, and can be embedded
relevance and trust in that information for use in climate services in a user’s risk landscape, taking account of socio-economic factors
(Box TS.11) and decision-making (high confidence). {1.2.3, 10.1.4, as well as physical climate changes. Storylines can also be used to
10.5, Cross-Chapter Box 10.3, 12.6} communicate climate information by narrative elements describing
and contextualizing the main climatological features and the relevant
The distillation process can vary substantially, as it needs to consider consequences in the user context and, as such, can be used as part
multiple lines of evidence on all physically plausible outcomes of a climate information distillation process. {1.4.4., Box 10.2, 11.2,
(especially when they are contrasting) relevant to a specific decision Box 11.2, Cross-Chapter Box 12.2}
required in response to a changing climate. Confidence in the
distilled regional climate information is enhanced when there is
agreement across multiple lines of evidence, so the outcome can
be limited if these are inconsistent or contradictory. For example, in
the Mediterranean region the agreement between different lines of
evidence, such as observations, projections by regional and global
models, and understanding of the underlying mechanisms, provides
high confidence in summer warming that exceeds the global average
(see Box TS.12). In a less clear-cut case for Cape Town, South Africa,
despite consistency among global model future projections, there
is medium confidence in a projected future drier climate due to the
lack of consistency in links between increasing greenhouse gases,
110
Technical Summary
Climate services involve providing climate information to assist decision-making, for example, about how extreme
rainfall will change to inform improvements in urban drainage. Since AR5, there has been a significant increase in the
range and diversity of climate service activities (very high confidence). The level of user-engagement, co-design and
co-production are factors determining the utility of climate services, while resource limitations for these activities
constrain their full potential. {12.6, Cross-Chapter Box 12.2}
Climate services include engagement from users and providers and an effective access mechanism; they are responsive to user needs
and based on integrating scientifically credible information and relevant expertise. Climate services are being developed across
regions, sectors, time scales and user-groups and include a range of knowledge brokerage and integration activities. These involve
identifying knowledge needs; compiling, translating and disseminating knowledge; coordinating networks and building capacity
through informed decision-making; analysis, evaluation and development of policy; and personal consultation.
Since AR5, climate change information produced in climate service contexts has increased significantly due to scientific and
technological advancements and growing user awareness, requirements and demand (very high confidence). Climate services are
growing rapidly and are highly diverse in their practices and products. The decision-making context, level of user engagement and co-
production between scientists, practitioners and intended users are important determinants of the type of climate service developed
and their utility for supporting adaptation, mitigation and risk management decisions. They require different types of user–producer
engagement depending on what the service aims to deliver (high confidence), and these fall into three broad categories: website-
based services, interactive group activities and focused relationships.
TS
Realization of the full potential of climate services is often hindered by limited resources for the co-design and co-production process,
including sustained engagement between scientists, service providers and users (high confidence). Further challenges relate to the
development and provision of climate services, generation of climate service products, communication with users, and evaluation of
their quality and socio-economic benefit. (Section TS.4.1) {1.2.3, 10.5.4, 12.6, Cross-Chapter Box 12.2, Glossary}
Box TS.12 | Multiple Lines of Evidence for Assessing Regional Climate Change and the
Interactive Atlas
A key novel element in the AR6 is the Working Group I Atlas, which includes the Interactive Atlas (https://interactive-
atlas.ipcc.ch/). The Interactive Atlas provides the ability to explore much of the observational and climate model data
used as lines of evidence in this assessment to generate regional climate information. {Atlas.2}
A significant innovation in the AR6 WGI Report is the Atlas. Part of its remit is to provide region-by-region assessment on changes in
mean climate and to link with other WGI chapters to generate climate change information for the regions. An important component
is the new online interactive tool, the Interactive Atlas, with flexible spatial and temporal analyses of much of the observed, simulated
past and projected future climate change data underpinning the WGI assessment. This includes the ability to generate global maps
and a number of regionally aggregated products (time series, scatter plots, tables, etc.) for a range of observations and ensemble
climate change projections of variables (such as changes in the climatic impact-drivers summarized in Table TS.5) from the Coupled
Model Intercomparison Project Phases 5 and 6 (CMIP5, CMIP6) and the Coordinated Regional Climate Downscaling Experiment
(CORDEX). The data can be displayed and summarized under a range of SSP-RCP scenarios and future time slices and also for different
global warming levels, relative to several different baseline periods. The maps and various statistics can be generated for annual mean
trends and changes or for any user-specified season. A new set of WGI reference regions is used for the regional summary statistics
and applied widely throughout the report (with the regions, along with aggregated datasets and the code to generate these, available
at the ATLAS GitHub: https://github.com/IPCC-WG1/Atlas).
Box TS.12, Figure 1 shows how the Interactive Atlas products, together with other lines of evidence, can be used to generate climate
information for an illustrative example of the Mediterranean summer warming. The lines of evidence include the understanding of
relevant mechanisms, dynamic and thermodynamic processes and the effect of aerosols in this case (Box TS.12, Figure 1a); trends
in observational datasets (which can have different spatial and temporal coverage; Box TS.12, Figure 1b, c); and attribution of these
trends and temperature projections from global and regional climate models at different resolutions, including single-model initial-
condition large ensembles (SMILEs; Box TS.12, Figure 1d, e). Taken together, this evidence shows there is high confidence that the
111
Technical Summary
projected Mediterranean summer temperature increase will be larger than the global mean, with consistent results from CMIP5 and
CMIP6 (Box TS.12, Figure 1e). However, CMIP6 results project both more pronounced warming than CMIP5 for a given emissions
scenario and time period and a greater range of changes (Box TS.12, Figure 1d). {10.6.4, Atlas.2, Atlas.8.4}
(a) Mechanisms of enhanced Mediterranean warming (b) Station locations (c) Temperature trend distribution
Past period (1960–2014)
OBS
Warm Atlantic Ocean Reduction of Surface drying Lapse rate difference Ensemble means
aerosols Over land stronger CMIP5
Enhanced warming than over sea CMIP6
sensible HighResMIP
Reduced Sea Land CORDEX EUR-44
heat flux
latent CORDEX EUR-11
heat flux MIROC6
Altitude
4
4 4
(°C)
3
CMIP6 (N=31)
2 2
2 SSP5-8.5
1
SSP3-7.0
0 0
SSP1-2.6 SSP2-4.5
SSP2-4.5 0 SSP1-2.6
–2 Long term SSP3-7.0
SSP5-8.5 0 1 2 3 4
2000 2020 2040 2060 2080 2100 Global temperature (°C)
Box TS.12, Figure 1 | Example of generating regional climate information from multiple lines of evidence for the case of Mediterranean
summer warming.
112
Technical Summary
TS.4.2 Drivers of Regional Climate Variability temperature is moderated or amplified by soil moisture feedback,
and Change snow/ice-albedo feedback, regional forcing from land-use/land-
cover changes, forcing from aerosol concentrations, or decadal/
Anthropogenic forcing, including GHGs and aerosols, but multi-decadal natural variability. Changes in local and remote
also regional land use and irrigation have all affected aerosol forcings lead to south–north gradients of the effective
observed regional climate changes (high confidence) and radiative forcing (hemispherical asymmetry). Along latitudes,
will continue to do so in the future (high confidence), with it is more uniform, with strong amplification of the temperature
various degrees of influence and response times, depending response towards the Arctic (medium confidence). The decrease
on warming levels, the nature of the forcing and the relative of SO2 emissions since the 1980s reduces the damping effect of
importance of internal variability. aerosols, leading to a faster increase in surface air temperature
that is most pronounced at mid- and high latitudes of the Northern TS
Since the late 19th century, major modes of variability Hemisphere, where the largest emissions reductions have taken
(MoVs) exhibited fluctuations in frequency and magnitude place (medium confidence). {1.3, 3.4.1, 6.3.4, 6.4.1, 6.4.3, 8.3.1,
at multi-decadal time scales, but no sustained trends 8.3.2, Box 8.1, 10.4.2, 10.6, 11.1.6, 11.3}
outside the range of internal variability (Table TS.4). An
exception is the Southern Annular Mode (SAM), which has Multi-decadal dimming and brightening trends in incoming solar
become systematically more positive (high confidence) and radiation at Earth’s surface occurred at widespread locations (high
is projected to be more positive in all seasons, except for confidence). Multi-decadal variation in anthropogenic aerosol
December–January–February (DJF), in high CO2 emissions emissions are thought to be a major contributor (medium confidence),
scenarios (high confidence). The influence of stratospheric but multi-decadal variability in cloudiness may also have played a
ozone forcing on the SAM trend has been reduced since the role. Volcanic eruptions affect regional climate through their spatially
early 2000s compared to earlier decades, contributing to heterogeneous effect on the radiative budget as well as through
the weakening of its positive trend as observed over 2000– triggering dynamical responses by favouring a given phase from
2019 (medium confidence). some MoVs, for instance. {1.4.1, Cross-Chapter Box 1.2, 2.2.1, 2.2.2,
3.7.1, 3.7.3, 4.3.1, 4.4.1, 4.4.4, Cross-Chapter Box 4.1, 7.2.2, 8.5.2,
In the near term, projected changes in most of the MoVs 10.1.4, 11.1.6, 11.3.1}
and related teleconnections will likely be dominated by
internal variability. In the long term, it is very likely that Historical urbanization affects the observed warming trends in cities
the precipitation variance related to El Niño–Southern and their surroundings (very high confidence). Future urbanization
Oscillation will increase. Physical climate storylines, will amplify the projected air temperature under different background
including the complex interplay between climate drivers, climates, with a strong effect on minimum temperatures that could
MoVs, and local and remote forcing, increase confidence be as large as the global warming signal (very high confidence)
in the understanding and use of observed and projected (Box TS.14). Irrigation and crop expansion have attenuated increases
regional changes. {2.4, 3.7, 4.3, 4.4, 4.5, 6.4, 8.3, 8.4, 10.3, in summer hot extremes in some regions, such as central North
10.4, 11.3} America (medium confidence) (Box TS.6). {Box 10.3, 11.1.6, 11.3}
113
Technical Summary
Since the late 19th century, major MoVs (Table TS.4) show contributed to observed temporal evolution in the Atlantic Multi-
no sustained trends, exhibiting fluctuations in frequency and decadal Variability (AMV) and associated regional teleconnections,
magnitude at multi-decadal time scales, except for the Southern especially since the 1960s, but there is low confidence in the
Annular Mode (SAM), which has become systematically more magnitude of this influence and the relative contributions of
positive (high confidence) (Table TS.4). It is very likely that human natural and anthropogenic forcings. Internal variability is the
influence has contributed to this trend from the 1970s to the main driver of Pacific Decadal Variability (PDV) observed since
1990s, and to the associated strengthening and southward shift the start of the instrumental records (high confidence), despite
of the Southern Hemispheric extratropical jet in austral summer. some modelling evidence for potential external influence. There is
The influence of stratospheric ozone forcing on the SAM trend has medium confidence that the AMV will undergo a shift towards a
been reduced since the early 2000s compared to earlier decades, negative phase in the near term. {2.4, 3.7.6, 3.7.7, 8.5.2, 4.4.3}
contributing to the weakening of its positive trend observed over
2000–2019 (medium confidence). By contrast, the cause of the
Northern Annular Mode (NAM) trend toward its positive phase
since the 1960s and associated northward shifts of Northern
Hemispheric extratropical jet and storm track in boreal winter
is not well understood. The evaluation of model performance on
simulating MoVs is assessed in Section TS.1.2.2. {2.3.3, 2.4, 3.3.3,
3.7.1, 3.7.2}
In the near term, the forced change in SAM in austral summer is likely
to be weaker than observed during the late 20th century under all
five SSPs assessed. This is because of the opposing influence in the
TS near to mid-term from stratospheric ozone recovery and increases in
other greenhouse gases on the Southern Hemisphere summertime
mid-latitude circulation (high confidence). In the near term, forced
changes in the SAM in austral summer are therefore likely to be
smaller than changes due to natural internal variability. In the long
term (2081–2100) under the SSP5-8.5 scenario, the SAM index is
likely to increase in all seasons relative to 1995–2014. The CMIP6
multi-model ensemble projects a long-term (2081–2100) increase
in the boreal wintertime NAM index under SSP3-7.0 and SSP5-8.5,
but regional associated changes may deviate from a simple shift
in the mid-latitude circulation due to a modified teleconnection
resulting from interaction with a modified mean background state.
{4.3.3, 4.4.3, 4.5.1, 4.5.3, 8.4.2}
114
Technical Summary
Table TS.4 | Summary of the assessments on modes of variability (MoVs) and associated teleconnections. (a) Assessments on observed changes since the start
of instrumental records, Coupled Model Intercomparison Project Phases 5 and 6 (CMIP5 and CMIP6) model performance, human influence on the observed changes, and
near-term (2021–2040) and mid- to long-term (2041–2100) changes. Curves schematically illustrate the assessed overall changes, with the horizontal axis indicating time,
and are not intended to precisely represent the time evolution. (b) Fraction of surface air temperature (SAT) and precipitation (pr) variance explained at interannual time scale
by each MoV for each AR6 region (numbers in each cell; in percent). Values correspond to the average of significant explained variance fractions based on HadCRUT, GISTEMP,
BerkeleyEarth and CRU-TS (for SAT) and GPCC and CRU-TS (for precipitation). Significance is tested based on F-statistics at the 95% level confidence, and a slash indicates that
the value is not significant in more than half of the available data sets. The colour scale corresponds to the sign and values of the explained variance as shown at the bottom. The
corresponding anomaly maps are shown in Annex IV. DJF: December–January–February. MAM: March–April–May. JJA: June–July–August. SON: September–October–November.
In (b), Northern Annular Mode (NAM) and El Niño–Southern Oscillation (ENSO) teleconnections are evaluated for 1959–2019, Southern Annular Mode (SAM) for 1979–2019,
Indian Ocean Basin (IOB), Indian Ocean Dipole (IOD), Atlantic Zonal Mode (AZM) and Atlantic Meridional Mode (AMM) for 1958–2019, and Pacific Decadal Variability (PDV)
and Atlantic Multi-decadal Variability (AMV) for 1900–2019. All data are linearly detrended prior to computation. (Section TS.1.2.2) {2.4, 3.7, 4.3.3, 4.4.3, 4.5.3, Table Atlas.1,
Annex IV}
(a) Assessments on MoV.
Contributed TS
through GHG Contributed
No robust Low No robust No robust No robust
Human influence on (all seasons) Not detected Not detected through
evidence agreement evidence evidence evidence
the observed changes & ozone aerosols
(DJF)
+
Internal Internal
No robust No robust No robust No robust Limited –
Near-term future variability variability
evidence evidence evidence evidence evidence Phase shift
changes (2021–2040) dominates All seasons dominates
except DJF from + to –
115
Technical Summary
Table TS.4 (continued): (b) Regional climate anomalies associated with MoV.
Mode NAM SAM ENSO IOB IOD AZM AMM PDV AMV
Season DJF DJF DJF MAM SON JJA JJA annual annual
Variable SAT pr SAT pr SAT pr SAT pr SAT pr SAT pr SAT pr SAT pr SAT pr
Mediterranean 28 58 7 19
Sahara 58 14 10 19 12 9 12 25
Western Africa 25 15 45 21 10 6 6 23
Central Africa 19 8 10 14 50 13 10 14 11
Africa
Tibetan Plateau 15 15 7 11 6 5 9
East Asia 7 20 23 9 9 13
South Asia 9 12 8 8 5
South East Asia 39 31 73 6 48 5 12 7
TS
Arabian Peninsula 32 10 24 20 5 13 7
Northern Australia 21 13 38 19 7 7 7
Australasia
Central Australia 14 21 12 18 22 20 7 7 6 5
East Australia 22 20 11 18 9 8 7 7 8
Southern Australia 11 23 40 8 3
New Zealand 16
Southern Central America 21 16 33 10 11 17 6 6 7
Central & South America
Pacific
Greenland/Iceland 42 8 7 44
Polar Terrestrial
Regions
Russian Arctic 25 10 6 11 8
West Antarctica 8 21
East Antarctica 38
colderwarmer drierwetter
Not significant in >50% of available data sets
40 30 20 0 20 30 40 40 30 20 0 20 30 40 Data unavailable in >50% of data sets
Temperature anomalies and Precipitation anomalies and
explained variance (%) explained variance (%)
116
Technical Summary
TS.4.2.3 Interplay Between Drivers of Climate Variability annual to decadal time scales (high confidence). The anthropogenic
and Change at Regional Scales signal in regional sea level change will emerge in most regions by
2100 (medium confidence). {9.2.4, 9.6.1, 10.4.1, 10.4.2, 10.4.3}
Anthropogenic forcing has been a major driver of regional mean
temperature change since 1950 in many sub-continental regions of Regional climate change is subject to the complex interplay between
the world (virtually certain). At regional scales, internal variability multiple external forcings and internal variability. Time evolution
is stronger, and uncertainties in observations, models and external of mechanisms operating at different time scales can modify the
forcing are all larger than at the global scale, hindering a robust amplitude of the regional-scale response of temperature, and
assessment of the relative contributions of greenhouse gases, both the amplitude and sign of the response of precipitation, to
stratospheric ozone, and different aerosol species in most of the cases. anthropogenic forcing (high confidence). These mechanisms include
Multiple lines of evidence, combining multi-model ensemble global non-linear temperature, precipitation and soil moisture feedbacks;
projections with those coming from single-model initial-condition slow and fast responses of SST patterns; and atmospheric circulation
large ensembles, show that internal variability is largely contributing changes to increasing GHGs. Land-use and aerosol forcings and
to the delayed or absent emergence of the anthropogenic signal in land–atmosphere feedback play important roles in modulating
long-term regional mean precipitation changes (high confidence). regional changes, for instance in weather and climate extremes (high
Internal variability in ocean dynamics dominates regional patterns on confidence). These can also lead to a higher warming of extreme
Pathway to understanding past and assessing future climate changes at regional scale
The South-Eastern South America (SES) case study
(a) Identification of climate drivers and phenomena for interpreting (b) Models simulations/evaluation of SES DJF precipitation over
SES observed precipitation trend and variability in austral summer (DJF) the historical period and 21st century based on 7 large ensembles
TS
AMV
Had
ley C
ell
Figure TS.21 | Example of the interplay between drivers of climate variability and change at regional scale to understand past and projected changes. The
figure intent is to show an illustrative pathway for understanding past, and anticipating future, climate change at regional scale in the presence of uncertainties. (a) Identification
of the climate drivers and their influences on climate phenomena contributing through teleconnection to South-Eastern South America (SES) summer (December–January–
February; DJF) precipitation variability and trends observed over 1950–2014. Drivers (red squares) include modes of variability as well as external forcing. Observed precipitation
linear trend from GPCC is shown on continents (green-brown colour bar in mm month–1 per decade) and the SES AR6 WGI reference region is outlined with the thick black
contour. Climate phenomena leading to local effects on SES are schematically presented (blue ovals). (b) Time series of decadal precipitation anomalies for DJF SES simulated
from seven large ensembles of historical plus RCP8.5 simulations over 1950–2100. Shading corresponds to the 5–95th range of climate outcomes given from each large
ensemble for precipitation (in mm month –1) and thick coloured lines stand for their respective ensemble mean. The thick time series in white corresponds to the multi-model
multi-member ensemble mean, with model contribution being weighted according to their ensemble size. GPCC observation is shown in the light black line with squares over
1950–2014, and the 1995–2014 baseline period has been retained for calculation of anomalies in all datasets. (c) Quantification of the respective weight (in percent) between
the individual sources of uncertainties (internal in grey, model in magenta and scenario in green) at near-term, mid-term and long-term temporal windows defined in AR6 and
highlighted in (b) for SES DJF precipitation. All computations are done with respect to 1995–2014, taken as the reference period, and the scenario uncertainty is estimated
from Coupled Model Intercomparison Project Phase 5 (CMIP5) using the same set of models as for the large ensembles that have run different Representative Concentration
Pathway (RCP) scenarios. {Figure 10.12a}
117
Technical Summary
temperatures compared to mean temperature (high confidence), which are relevant for the region. In fact, local changes over SES
and possibly cooling in some regions (medium confidence). The in terms of moisture convergence, ascending motion and storm-
soil moisture–temperature feedback was shown to be relevant for track locations depend on these climate phenomena, and they are
past and present-day heatwaves based on observations and model overall responsible for the observed precipitation trends. Projections
simulations. {10.4.3, 11.1.6, 11.3.1} suggest continuing positive trends in rainfall over SES in the near-
term in response to GHG emissions scenarios. Multi-model mean and
South-Eastern South America (SES) is one of the AR6 WGI reference ensemble spread are not sufficient to characterize situations where
regions (outlined with black thick contour in Figure TS.21a), and different models simulate substantially different or even opposite
it is used here as an illustrative example of the interplay between changes (high confidence). In such cases, physical climate storylines
drivers of climate variability and change at regional scale. Austral addressing possible outcomes for climate phenomena shown to
summer (DJF) precipitation positive trends have been observed play a role in the variability of the region of interest can aid the
over the region during 1950–2014. Drivers of this change include interpretation of projection uncertainties. In addition, single-model
MoVs, such as AMV, ENSO, and PDV, as well as external forcing, initial-condition large ensembles of many realizations of internal
like GHG increases and ozone depletion together with aerosols (as variability are required to separate internal variability from forced
illustrated in Figure TS.21a). Modes of variability and external forcing changes (high confidence) and to partition the different sources of
collectively affect climate phenomena, such as the Hadley cell width uncertainties as a function of future assessed periods. {10.3.4, 10.4.2,
and strength, Rossby waves activity emerging from the large-scale Figure 10.12a}
tropical SST anomalies, and the Southern Hemisphere polar vortex,
TS Global land monsoon precipitation decreased from the 1950s to the 1980s, partly due to anthropogenic aerosols, but
has increased since then in response to GHG forcing and large-scale multi-decadal variability (medium confidence).
Northern Hemispheric anthropogenic aerosols weakened the regional monsoon circulations in South Asia, East Asia
and West Africa during the second half of the 20th century, thereby offsetting the expected strengthening of monsoon
precipitation in response to GHG-induced warming (high confidence).
During the 21st century, global land monsoon precipitation is projected to increase in response to GHG warming in
all time horizons and scenarios (high confidence). Over South and South East Asia, East Asia and the central Sahel,
monsoon precipitation is projected to increase, whereas over North America and the far western Sahel it is projected
to decrease (medium confidence). There is low confidence in projected precipitation changes in the South American
and Australian-Maritime Continent monsoons. At global and regional scales, near-term monsoon changes will be
dominated by the effects of internal variability (medium confidence). {2.3, Cross-Chapter Box 2.4, 3.3, 4.4, 4.5, 8.2,
8.3, 8.4, 8.5, Box 8.1, Box 8.2, 10.6}
Global Monsoon
Paleoclimate records indicate that during warm climates, like the mid-Pliocene Warm Period, monsoon systems were stronger
(medium confidence). In the instrumental records, global summer monsoon precipitation intensity has likely increased since the
1980s, dominated by Northern Hemisphere summer trends and large multi-decadal variability. Contrary to the expected increase of
precipitation under global warming, the Northern Hemisphere monsoon regions experienced declining precipitation from the 1950s
to 1980s, which is partly attributable to the influence of anthropogenic aerosols (medium confidence) (Box TS.13, Figure 1). {2.3.1,
Cross-Chapter Box 2.4, 3.3.2, 3.3.3}
With continued global warming, it is likely that global land monsoon precipitation will increase during this century (Box TS.13, Figure 1),
particularly in the Northern Hemisphere, although the monsoon circulation is projected to weaken. A slowdown of the tropical circulation
with global warming can partly offset the warming-induced strengthening of precipitation in monsoon regions (high confidence). In the
near term, global monsoon changes are likely to be dominated by the effects of internal variability and model uncertainties (medium
confidence). In the long term, global monsoon rainfall change will feature a robust north–south asymmetry characterized by a greater
increase in the Northern Hemisphere than in the Southern Hemisphere and an east–west asymmetry characterized by enhanced Asian–
African monsoons and a weakened North American monsoon (medium confidence). {4.4.1, 4.5.1, 8.4.1}
Regional Monsoons
Paleoclimate reconstructions indicate stronger monsoons in the Northern Hemisphere but weaker ones in the Southern Hemisphere
during warm periods, particularly for the South and South East Asian, East Asian, and North and South American monsoons, with
the opposite occurring during cold periods (medium confidence). It is very likely that Northern Hemispheric anthropogenic aerosols
weakened the regional monsoon circulations in South Asia, East Asia and West Africa during the second half of the 20th century,
118
Technical Summary
NAmerM SAsia M
EAsia M
WAfriM
EqAmer
SAfri
SAmer M AusMCM
0 x
x
xx
x
x
x x
–0.04 x
x
–0.08 x
x
–0.12
NAmerM WA fri M SAsiaM EAsiaM EqAmer SAmerM SAfri AusMCM GM
10
0
Long term
Near term
Mid term
–10
–20
NAmerM WA fri M SAsiaM EAsiaM EqAmer SAmerM SAfri AusMCM GM
Box TS.13, Figure 1 | Global and regional monsoons: past trends and projected changes. The intent of this figure is to show changes in precipitation over
regional monsoon domains in terms of observed past trends, how greenhouse gases and aerosols relate to these changes, and in terms of future projections in one
intermediate emissions scenario in the near, medium and long term. (a) Global (black contour) and regional monsoons (colour shaded) domains. The global monsoon
(GM) is defined as the area with local summer-minus-winter precipitation rate exceeding 2.5 mm day–1 (see Annex V). The regional monsoon domains are defined
based on published literature and expert judgement (see Annex V) and accounting for the fact that the climatological summer monsoon rainy season varies across
the individual regions. Assessed regional monsoons are South and South East Asia (SAsiaM, Jun–July–August–September), East Asia (EAsiaM, June–July–August),
West Africa (WAfriM, June–July–August–September), North America (NAmerM, July–August–-September), South America (SAmerM, December–January–February),
Australia and Maritime Continent Monsoon (AusMCM, December–January–February). Equatorial South America (EqSAmer) and South Africa (SAfri) regions are also
shown, as they receive unimodal summer seasonal rainfall although their qualification as monsoons is subject to discussion. (b) Global and regional monsoons
precipitation trends based on DAMIP CMIP6 simulations with both natural and anthropogenic (ALL), greenhouse gas only (GHG), aerosols only (AER) and natural
only (NAT) radiative forcing. Weighted ensemble means are based on nine Coupled model Intercomparison Project Phase 6 (CMIP6) models contributing to the MIP
(with at least three members). Observed trends computed from CRU, GPCP and APHRO (only for SAsiaM and EAsiaM) datasets are shown as well. (c) Percentage
change in projected seasonal mean precipitation over global and regional monsoons domain in the near term (2021–2040), mid-term (2041–2060), and long term
(2081–2100) under SSP2-4.5 based on 24 CMIP6 models. {Figures 8.11 and 8.22}
119
Technical Summary
thereby offsetting the expected strengthening of monsoon precipitation in response to GHG-induced warming (Box TS.13, Figure 1).
Multiple lines of evidence explain this contrast over South Asia, with the observed trends dominated by the effects of aerosols, while
future projections are mostly driven by GHG increases. The recent partial recovery and enhanced intensity of monsoon precipitation
over West Africa is related to the growing influence of GHGs with an additional contribution due to the reduced cooling effect of
anthropogenic aerosols, emitted largely from North America and Europe (medium confidence). For other regional monsoons, that
is, North and South America and Australia, there is low confidence in the attribution of recent changes in precipitation (Box TS.13,
Figure 1) and winds. {2.3.1, 8.3.1, 8.3.2, Box 8.1, 10.6.3}
Projections of regional monsoons during the 21st century indicate contrasting (region-dependent) and uncertain precipitation and
circulation changes. The annual contrast between the wettest and driest month of the year is likely to increase by 3–5% per degree
Celsius in most monsoon regions in terms of precipitation, precipitation minus evaporation, and runoff (medium confidence). For the
North American monsoon, projections indicate a decrease in precipitation, whereas increased monsoon rainfall is projected over South
and South East Asia and over East Asia (medium confidence) (Box TS.13, Figure 1). West African monsoon precipitation is projected
to increase over the central Sahel and decrease over the far western Sahel (medium confidence). There is low confidence in projected
precipitation changes in the South American and Australian-Maritime Continent regional monsoons (for both magnitude and sign)
(Box TS.13, Figure 1). There is medium confidence that the monsoon season will be delayed in the Sahel and high confidence that it
will be delayed in North and South America. {8.2.2, 8.4.2.4, Box 8.2}
Overall, long-term (2081–2100) future changes in regional monsoons like the South and South East Asian monsoon are generally
consistent across global (including high-resolution) and regional climate models and are supported by theoretical arguments.
Uncertainties in simulating the observed characteristics of regional monsoon precipitation are related to varying complexities of
regional monsoon processes and their responses to external forcing, internal variability, and deficiencies in representing monsoon warm
rain processes, organized tropical convection, heavy orographic rainfall and cloud–aerosol interactions. {8.3.2, 8.5.1, 10.3.3, 10.6.3}
TS.4.3 Regional Climate Change and Implications for Relative sea level rise is very likely to virtually certain
Climate Extremes and Climatic Impact-Drivers (depending on the region) to continue during the 21st
century, contributing to increased coastal flooding in low-
Current climate in all regions is already distinct from the lying areas (high confidence) and coastal erosion along
climate of the early or mid-20th century with respect to most sandy coasts (high confidence). Sea level will continue
several climatic impact-drivers (CIDs), resulting in shifting to rise beyond 2100 (high confidence) (Box TS.4).
magnitude, frequency, duration, seasonality and spatial
extent of associated climate indices (high confidence). Every region of the world will experience concurrent
It is very likely that mean temperatures have increased changes in multiple CIDs by mid-century or at 2°C global
in all land regions and will continue to increase at rates warming and above (high confidence). Even for the current
greater than the global average (high confidence). The climate, climate change-induced shifts in CID distributions
frequency of heat and cold extremes have increased and and event probabilities, some of which have occurred over
decreased, respectively. These changes are attributed to recent decades, are relevant for risk assessments. {11.9,
human influence in almost all regions (medium to high 12.1, 12.2, 12.4, 12.5, Atlas.3–Atlas.11}
confidence) and will continue through the 21st century
(high confidence). In particular, extreme heat would exceed An overview of changes in regional CIDs (introduced in Section TS.1)
critical thresholds for health, agriculture and other sectors is given in Table TS.5, which summarizes multiple lines of evidence
more frequently by the mid 21st century with 2°C of global on regional climate change derived from observed trends, attribution
warming (high confidence). of these trends and future projections. The level of confidence and
120
Technical Summary
TS
121
TS
Table TS.5 | Summary of confidence for climatic impact-driver changes in each AR6 WGI reference region (illustrated in Figure TS.25) across multiple lines of evidence for observed, attributed and projected
Technical Summary
122
directional changes. The colours represent their projected aggregate characteristic changes for the mid-21st century, considering scenarios RCP4.5, SSP2-4.5, SRES A1B, or above (RCP6.0, RCP8.5, SSP3-7.0, SSP5-8.5, SRES A2), which
approximately encompasses global warming levels of 2.0°C to 2.4°C. Arrows indicate medium to high confidence trends derived from observations, and asterisks indicate medium and high confidence in attribution of observed changes. (North
Africa is not an AR6 WGI reference region, but assessment here is based upon the African portion of the Mediterranean reference region). {Tables 12.3–12.11 and Tables 11.4–11.21}
Climatic Impact-driver
Mean air temperature Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Africa
Sahara ↗ ↗ ** ↘ ** ↗ 4 ↗ ↗ ↗
Western Africa ↗ ↗ ** ↘ ** 1 ↗ ↗1 ↗1 ↗1 ↗ 4 ↗ ↗ ↗
Madagascar ↗ ↗ ↘ 3 ↗ 4,5 ↗ ↗ ↗
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
Table TS.5 (continued)
Climatic Impact-driver
Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Asia
Technical Summary
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
123
TS
TS
Table TS.5 (continued)
Technical Summary
124
Climatic Impact-driver
Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Australasia
New Zealand ↗ ↗ ↘ ** 2 4 8 ↘6 ↗ 7 ↗ ↗ ↗
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
Table TS.5 (continued)
Climatic Impact-driver
Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Mean air temperature
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Central and South America
Southern Central
America ↗ ↗ ** ↘ ** 2 ↗ 3 ↗ ↗ ↗
North-Western
South America ↗ ↗ *** ↘ *** ↗ 3,4 ↗ ↗ ↗
Northern South
America ↗ ↗ ** ↘ ** 2 ↗ 3,4 ↗ ↗ ↗
South American
Monsoon ↗ ↗ ** ↘ ** ↗1 ↗
North-Eastern
South America ↗ ↗ ** ↘ ** ↘ ↗ ↗ 3,4 ↗ ↗ ↗
South-Western
South America ↗ ↗ ** ↘ ** ↘ ↗ ↗ 3 ↗ ↗ ↗
South-Eastern
South America ↗ ↗ *** ↘ *** ↘ ↗ ↗ ↘ ↗ 3 ↗ ↗ ↗
Southern South
America ↗ ↘ ↗ 3 ↗ ↗ ↗
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
Technical Summary
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
125
TS
TS
Table TS.5 (continued)
Technical Summary
126
Climatic Impact-driver
Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Europe
North America
North Central
America ↗ ↗ ** ↘ ** ↗ ↗ ↗ ↗ 2 ↗ ↗ ↗
Western North 6,7
America ↗ ↗ ** ↘ ** 3 5 5 4,7 ↗ ** ↗ 6.7 ↘ 8 ↗6 ↘1 ↘ 1 1 ↗ ↗5 2 ↗ ↗ ↗
Central North
America ↗ ↗ ↗ ** 7 7 7 ↘ 8 4 ↘ ↘ ↗ ↗ 2 ↗ ↗ ↗
Eastern North
America ↗ ↗5 ↗ 7 ↘ 8 ↘1 ↘ 1 1 ↗ ↗ 2 ↗ ↗ ↗
North-Eastern
North America ↗ ↗ *** ↘ *** ↘ 5 5 6,7 6,7 8 ↘ 1,6 ↘ ↘ 1 4 ↗ 4,6 2,6 ↗ ↗ ↗
North-Western
North America ↗ ↗ *** ↘ *** ↘ 5 6 5 6,7 ↗ 6,7 8 ↘1 ↘ ↘ 1,6 ↗9 ↗ 2 ↗ ↗ ↗
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
Table TS.5 (continued)
Climatic Impact-driver
Heat and Cold Wet and Dry Wind Snow and Ice Coastal and Oceanic Other
Extreme heat
Cold spell
Frost
Mean precipitation
River flood
pluvial flood
Heavy precipitation and
Landslide
Aridity
Hydrological drought
Fire weather
Tropical cyclone
Permafrost
Hail
Snow avalanche
Coastal flood
Coastal erosion
Marine heatwave
Radiation at surface
Small Islands
Caribbean ↗ ↗ ** 5 ↗ 6 ↗ ↗ ↗
Pacific ↗ ↗ **1 2 3 4 5 ↗ 6 ↗ ↗ ↗
Greenland and
Iceland ↗ ↗ ** ↘ ** ↗ ↗ ↘3 2,3 ↘1 ↘ ↘ ↗5 ↗ ↗
Arctic North
Europe ↗ ↗ ↗ ↘3 2,3 ↘1 ↘ ↘ ↗6 7 ↗ ↗
East Antarctica ↗ ↗
Technical Summary
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
Key for level of confidence in future changes High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
127
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
TS
TS
Table TS.5 (continued)
Technical Summary
128
Climatic Impact-driver
Marine heatwave
Ocean acidity
Ocean salinity
Dissolved oxygen
Sea ice
Oceans
Arabian Sea ↗ ↗ ↗
Bay of Bengal ↗ ↗ ↗
Southern Ocean
Note: There are several region-specific qualifiers/exceptions attached to some of the directions of change/confidence levels indicated above. {12.4}
Key for observational trend evidence ↗ Past upward trend (medium or higher confidence) ↘ Past downward trend (medium or higher confidence)
Key for attribution evidence *** High confidence (or more) ** Medium confidence
High confidence Medium confidence Low confidence Medium confidence High confidence Not broadly
Key for level of confidence in future changes
of increase (or more) of increase (or more) in direction of change of decrease of decrease relevant
Technical Summary
Notes:
Africa (projections)
1. Contrasted regional signal: drying in western portions and wetting in eastern portions
2. Likely increase over the Ethiopian Highlands
3. Medium confidence of decrease in frequency and increase in intensity
4. Along sandy coasts and in the absence of sufficient sediment supply from terrestrial or offshore sources
5. Substantial parts of the East Southern Africa and Madagascar coast are projected to prograde if present-day ambient shoreline change rates continue
Asia (projections)
1. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat.
2. Substantial parts of the coasts in these regions are projected to prograde if present-day ambient shoreline change rates continue
3. Tropical cyclones decrease in number but increase in intensity
4. High confidence of decrease in Indonesia (Atlas.5.4.5)
5. Medium confidence of decreasing in summer and increasing in winter
Australasia (projections)
1. High confidence of decrease in the south-west of the state of Western Australia
2. Medium confidence of decrease in north and east and increase in south and west
3. High confidence of increase in the south-west of the state of Western Australia
4. Medium confidence of increase in the north and east and decrease in south and west
5. Low confidence of increasing intensity, and high confidence of decreasing occurrence
6. High confidence of decrease in glacier volume, medium confidence of decrease in snow
7. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat
Central and South America (projections)
1. Increase in extreme flow in the Amazon basin
2. Tropical cyclones decrease in number but increase in intensity
3. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat. TS
4. Substantial parts of the North-Western South America, Northern South America and North-Eastern South America coasts are projected to prograde if present-day
ambient shoreline change rates continue
Europe (projections)
1. Excluding southern United Kingdom
2. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat
3. The Baltic Sea shoreline is projected prograde if present-day ambient shoreline change rates continue.
4. For the Alps, conditions conducive to landslides are expected to increase
5. Low confidence of decrease in the southernmost part of the region
6. General decrease except in Aegean Sea
7. Medium confidence of decrease in frequency and increase in intensities
8. Except in the Northern Baltic Sea region
North America (projections)
1. Snow may increase in some high elevations and during the cold season and decrease in other seasons and at lower elevations
2. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat.
3. Increasing in northern regions and decreasing toward the south
4. Decreasing in northern regions and increasing toward the south
5. Higher confidence in northern regions and lower toward the south
6. Higher confidence in southern regions and lower toward the north
7. Higher confidence in increase for some climatic impact-driver indices during summertime
8. Increase in convective conditions but decrease in winter extratropical cyclones
9. Relative sea level rise reduced given land uplift in Southern Alaska
Small Islands (projections)
1. Very high confidence in the direction of change, but low to medium confidence in the magnitude of change due to model uncertainty
2. Decrease in eastern Pacific and southern Pacific subtropics, but increase in parts of western and equatorial Pacific; with seasonal variation in future changes
3. High confidence in increase in extreme rain frequency and intensity in western tropical Pacific; low confidence in magnitude of change due to model bias
4. Increase in southern Pacific
5. Increase in intensity; decrease in frequency except over central North Pacific.
6. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat.
Polar Terrestrial Regions (projections)
1. Snow may increase in some high elevations and during the cold season and decrease in other seasons and at lower elevations
2. Higher confidence in southern regions and lower toward the north
3. Higher confidence in increase for some climatic impact-driver indices during summertime
4. Glaciers decline even as some regional snow climatic impact-driver indices increase
5. Decreasing in west and increasing in east
6. Except for Northern Baltic Sea coasts where relative sea levels fall
7. Along sandy coasts and in the absence of additional sediment sinks/sources or any physical barriers to shoreline retreat
129
Technical Summary
Heat
increase in mean
temperature, extreme heat
CIDs changing in all the
regions of the cluster with
high confidence
Cold
decrease in cold spell, frost
Mean Mean
precipitation precipitation
decrease increase
River Pluvial
flooding flooding
increase increase
(3) (4) (5)
Hotter, less snow and ice and Hotter and less snow/ice Hotter
more pluvial flooding and in some regions more and in some regions more
and in some regions more pluvial flooding or river pluvial flooding or mean Fire
precipitation or fire weather or flooding or mean precipita- precipitation or both weather
both tion or both increase
130
Technical Summary
(b) Number of land & coastal regions (i) and open-ocean regions (ii) where each climatic impact-driver (CID) is projected to
increase or decrease with high confidence (dark shade) or medium confidence (light shade)
(i) (ii)
Heat & Cold Wet & Dry
Heavy precipitation and pluvial flood Wind Snow & Ice Other Coastal Open Ocean
Radiation at surface
Severe wind storm
Mean precipitation
Dissolved oxygen
Marine heatwave
Marine heatwave
Snow avalanche
Tropical cyclone
Coastal erosion
Ocean salinity
Ocean acidity
Ocean acidity
Extreme heat
Coastal flood
NUMBER OF LAND
Fire weather
Permafrost
River flood
Cold spell
Landslide
Aridity
Frost
Hail
55
45 TS
35
25
15 15
5 5
5 5
15 15
25
35
45
55
Climatic impact-drivers (CIDs) are physical climate system conditions (e.g., means, events, extremes) that affect an element of
society or ecosystems. Depending on system tolerance, CIDs and their changes can be detrimental, beneficial, neutral, or a
mixture of each across interacting system elements and regions. The CIDs are grouped into seven types, which are summarized
under the icons in sub-panels (i) and (ii). All regions are projected to experience changes in at least 5 CIDs. Almost all (96%) are
projected to experience changes in at least 10 CIDs and half in at least 15 CIDs. For many CID changes, there is wide
geographical variation, and so each region is projected to experience a specific set of CID changes. Each bar in the chart interactive-atlas.ipcc.ch
represents a specific geographical set of changes that can be explored in the WGI Interactive Atlas.
Figure TS.22 | Synthesis of the geographical distribution of climatic impact-drivers changes and the number of AR6 WGI reference regions where they
are projected to change.
131
Technical Summary
Figure TS.22 (continued): Panel (a) shows the geographical location of regions belonging to one of five groups characterized by a specific combination of changing climatic
impact-drivers (CIDs). The five groups are represented by the five different colours, and the CID combinations associated with each group are represented in the corresponding
‘fingerprint’ and text below the map. Each fingerprint comprises a set of CIDs projected to change with high confidence in every region in the group and a second set of CIDs,
one or more of which are projected to change in each region with high or medium confidence. The CID combinations follow a progression from those becoming hotter and drier
(group 1) to those becoming hotter and wetter (group 5). In between (groups 2–4), the CIDs that change include some becoming drier and some wetter and always include
a set of CIDs which are getting hotter. Tropical cyclones and severe wind CID changes are represented on the map with black dots in the regions affected. Regions affected
by coastal CID changes are described by text on the map. The five groups are chosen to provide a reasonable level of detail for each region while not overwhelming the map
with a full summary of all aspects of the assessment, which is available in Table TS.5 and can be visualized in the Regional Synthesis component of the Interactive Atlas. The
CID changes summarized in the figure represent high and medium confidence changes for the mid-21st century, considering scenarios SSP2-4.5, RCP4.5, SRES A1B, or above
(SSP3-7.0, SSP5-8.5, RCP6.0, RCP8.5, SRES A2), which approximately encompasses global warming levels of 2.0°C to 2.4°C.
The bar chart in panel (b) shows the numbers of regions where each CID is increasing or decreasing with medium or high confidence for all land regions and ocean regions
listed in Table TS.5. The colours represent the direction of change and the level of confidence in the change: purple indicates an increase while brown indicates a decrease; darker
and lighter shades refer to high and medium confidence, respectively. Lighter background colours represent the maximum number of regions for which each CID is broadly
relevant. Sub-panel (i) shows the 30 CIDs relevant to the land and coastal regions while sub-panel (ii) shows the 5 CIDs relevant to the open ocean regions. Marine heatwaves
and ocean acidity are assessed for coastal ocean regions in panel (i) and for open ocean regions in panel (ii). Changes refer to a 20- to 30-year period centred around 2050
and/or consistent with 2°C global warming compared to a similar period within 1960–2014, except for hydrological drought and agricultural and ecological drought, which
is compared to 1850–1900. Definitions of the regions are provided in Atlas.1, the Interactive Atlas (https://interactive-atlas.ipcc.ch/) and Chapter 12. (Table TS.5, Figure TS.24)
{11.9, 12.2, 12.4, Atlas.1}
TS.4.3.1 Common Regional Changes in Climatic Impact-Drivers regions (high confidence). In many tropical regions, the number of
days per year where a heat index of 41°C is exceeded would increase
Heat and cold: Changes in temperature-related CIDs such as mean by more than 100 days relative to the recent past under SSP5-8.5,
temperatures, growing season length, and extreme heat and frost have while this increase will be limited to less than 50 days under SSP1-2.6
already occurred (high confidence), and many of these changes have (high confidence) (Figure TS.6). The number of days per year where
been attributed to human activities (medium confidence). Over all temperature exceeds 35°C would increase by more than 150 days in
TS land regions with sufficient data (i.e., all except Antarctica), observed many tropical areas, such as the Amazon basin and South East Asia,
changes in temperature have already clearly emerged outside the by the end of century for the SSP5-8.5 scenario, while it is expected to
range of internal variability, relative to 1850–1900 (Figure TS.23). In increase by less than 60 days in these areas under SSP1-2.6 (except
tropical regions, recent past temperature distributions have already for the Amazon Basin) (high confidence) (Figure TS.24). {4.6.1, 11.3,
shifted to a range different to that of the early 20th century (high 11.9, 12.4, 12.5.2, Atlas}
confidence) (Section TS.1.2.4). Most land areas have very likely
warmed by at least 0.1°C per decade since 1960, and faster in recent Wet and dry: Compared to the global scale, precipitation internal
decades. On regional-to-continental scales, trends of increased variability is stronger at the regional scale while uncertainties in
frequency of hot extremes and decreased frequency of cold extremes observations, models and external forcing are all larger. However,
are generally consistent with the global-scale trends in mean GHG forcing has driven increased contrasts in precipitation amounts
temperature (high confidence). In a few regions, trends are difficult between wet and dry seasons and weather regimes over tropical land
to assess due to limited data availability. {2.3.1.1, 11.3, 11.9, 12.4, areas (medium confidence), with a detectable precipitation increase
Atlas.3.1} in the northern high latitudes (high confidence) (Box TS.6). The
frequency and intensity of heavy precipitation events have increased
Warming trends observed in recent decades are projected to continue over a majority of land regions with good observational coverage
over the 21st century and over most land regions at a rate higher (high confidence). A majority of land areas have experienced
than the global average (high confidence). For given global warming decreases in available water in dry seasons due to human-induced
levels, model projections from CMIP6 show future regional warming climate change associated with changes in evapotranspiration
changes that are similar to those projected by CMIP5. However, (medium confidence). Global hydrological models project a larger
projected regional warming in CMIP6 for given time periods and fraction of land areas to be affected by an increase rather than by
emissions scenarios has a wider range with a higher upper limit a decrease in river floods (medium confidence). Extreme precipitation
compared to CMIP5 because of the higher climate sensitivity in some and pluvial flooding will increase in many regions around the world
CMIP6 models and differences in the forcings. {Atlas.3–Atlas.11} on almost all continents (high confidence), but regional changes
in river floods are more uncertain than changes in pluvial floods
Under RCP8.5/SSP5-8.5, it is likely that most land areas will experience because complex hydrological processes, including land cover and
further warming of at least 4°C compared to a 1995–2014 baseline human water management, are involved. {8.2.2.1, 8.3.1, Box 8.2,
by the end of the 21st century, and in some areas significantly more. 10.4.1, 11.5, 11.6, 11.9, 12.4, 12.5.1, Atlas.3.1}
At increasing warming levels, extreme heat will exceed critical
thresholds for health, agriculture and other sectors more frequently Wind: Mean wind speed has decreased over most land areas with
(high confidence), and it is likely that cold spells will become less good observational coverage (medium confidence). It is likely that
frequent towards the end of the century. For example, by the end the global proportion of major tropical cyclone (TC) intensities
of the 21st century, dangerous humid heat thresholds, such as the (Categories 3–5) over the past four decades has increased. The
National Oceanic and Atmospheric Administration (NOAA) heat index proportion of intense TCs, average peak TC wind speeds, and peak
(HI) threshold of 41°C, will be exceeded much more frequently under wind speeds of the most intense TCs will increase on the global scale
the SSP5-8.5 scenario than under SSP1-2.6 and will affect many with increasing global warming (high confidence). {11.7.1}
132
Technical Summary
TS
Before 1981 1981–1988 1989–1996 1997–2004 2005–2012 2013–2020
Dataset: CRUTEM5. Temperature changes relative to 1850–1900. Grey: not enough data.
Figure TS.23 | Time period during which the signals of temperature change in observed data aggregated over the reference regions emerged from the
noise of annual variability in the respective aggregated data, using a signal-to-noise ratio of two as the threshold for emergence. The intent of this figure
is to show, for the AR6 WGI reference regions, when a signal of annual mean surface temperature change emerged from the noise of annual variability in two global datasets
and thus also provide some information on observational uncertainty. Emergence time is calculated for two global observational datasets: (a) Berkeley Earth and (b) CRUTEM5.
Regions in the CRUTEM5 map are shaded grey when data are available over less than 50% of the area of the region. (Section TS.1.2.4) {Figure Atlas.11}
133
Technical Summary
(b) TX35 for 2081−2100 (RCP2.6) rel. to 1995−2014 (f) TX35 for 2081−2100 (SSP1-2.6) rel. to 1995−2014 (j) TX35 for 2081−2100 (SSP1-2.6) rel. to 1995−2014
(c) TX35 for 2041−2060 (RCP8.5) rel. to 1995−2014 (g) TX35 for 2041−2060 (SSP5-8.5) rel. to 1995−2014 (k) TX35 for 2041−2060 (SSP5-8.5) rel. to 1995−2014
(d) TX35 for 2081−2100 (RCP8.5) rel. to 1995−2014 (h) TX35 for 2081−2100 (SSP5-8.5) rel. to 1995−2014 (l) TX35 for 2081−2100 (SSP5-8.5) rel. to 1995−2014
TS
Figure TS.24 | Projected change in the mean number of days per year with maximum temperature exceeding 35°C for Coupled Model Intercomparison
Project Phase 5 (CMIP5; first column), Phase 6 (CMIP6; second column) and Coordinated Regional Climate Downscaling Experiment (CORDEX; third
column) ensembles. The intent of this figure is to show that there is a consistent message about the patterns of projected change in extreme daily temperatures from the
CMIP5, CMIP6 and CORDEX ensembles. The map shows the median change in the number of days per year between the mid-century (2041–2060) or end-century (2081–2100)
and historical (1995–2014) periods for the CMIP5 and CORDEX RCP8.5 and RCP2.6 scenario ensembles and the CMIP6 SSP5-8.5 and SSP1-2.6 scenario ensembles. Hatching
indicates areas where less than 80% of the models agree on the sign of change. {Interactive Atlas}
Snow and ice: Many aspects of the cryosphere either have seen Coastal and oceanic: There is high confidence that SST will
significant changes in the recent past or will see them during the increase in all oceanic regions except the North Atlantic. Regional
21st century (high confidence). Glaciers will continue to shrink sea level change has been the main driver of changes in extreme
and permafrost to thaw in all regions where they are present (high sea levels across the quasi-global tide gauge network over the
confidence). Also, it is virtually certain that snow cover will experience 20th century (high confidence). With the exception of a few regions
a decline over most land regions during the 21st century, in terms with substantial land uplift, relative sea level rise is very likely to
of water equivalent, extent and annual duration. There is high virtually certain (depending on the region) to continue during the
confidence that the global warming-induced earlier onset of spring 21st century, contributing to increased coastal flooding in low-lying
snowmelt and increased melting of glaciers have already contributed areas (high confidence) and coastal erosion along most sandy
to seasonal changes in streamflow in high-latitude and low-elevation coasts (high confidence) over the 21st century. In the open ocean,
mountain catchments. Nevertheless, it is very likely that some high- acidification, changes in sea ice, and deoxygenation have already
latitude regions will experience an increase in winter snow water emerged in many areas (high confidence). Marine heatwaves are
equivalent due to the effect of increased snowfall prevailing over also expected to increase around the globe over the 21st century
warming-induced increased snowmelt. (Section TS.2.5) {8.2.2.1, (high confidence). (Section TS.2.4) {Box 9.2, 9.2.1.1, 9.6, 9.6.4,
8.3.1, Box 8.2, 9.4, 9.5.1, 9.5.2, 12.4, Atlas.4–Atlas.9, Atlas.11} 9.6.4.2, 12.4}
134
Technical Summary
Other variables and concurrent CID changes: It is virtually certain In addition to the main changes summarized above and in
that atmospheric CO2 and oceanic pH will increase in all climate Section TS.4.3.1, additional details per CID are given below.
scenarios, until net zero CO2 emissions are achieved (Section TS.2.2).
In nearly all regions, there is low confidence in changes in hail, ice Heat and cold: Observed and projected increases in mean
storms, severe storms, dust storms, heavy snowfall, and avalanches, temperature and a shift toward heat extreme characteristics are
although this does not indicate that these CIDs will not be affected broadly similar to the generic pattern described in Section TS.4.3.1.
by climate change. For such CIDs, observations are often short- {2.3.1.1.2, 11.3, 11.9, 12.4.1.1, Atlas.4.2, Atlas.4.4}
term or lack homogeneity, and models often do not have sufficient
resolution or accurate parametrizations to adequately simulate them Wet and dry: Mean precipitation changes have been observed
over climate change time scales. The probability of compound events over Africa, but the historical trends are not spatially coherent (high
has increased in the past due to human-induced climate change and confidence). North Eastern Africa, East Southern Africa and Central
will likely continue to increase with further global warming, including Africa have experienced a decline in rainfall since about 1980 and
for concurrent heatwaves and droughts, compound flooding, and the parts of West Africa an increase (high confidence). Increases in the
possibility of connected sectors experiencing multiple regional extreme frequency and/or the intensity of heavy rainfall have been observed in
events at the same time (for example, in multiple breadbaskets) (high East and West Southern Africa, and the eastern Mediterranean region
confidence). {5.3.4.2, 11.8, Box 11.3, Box 11.4, 12.4} (medium confidence). Increasing trends in river flood occurrence can
be identified beyond 1980 in East and West Southern Africa (medium
TS.4.3.2 Region-by-Region Changes in Climatic Impact-Drivers confidence) and Western Africa (high confidence). However, Northern
Africa and West Southern Africa are likely to have a reduction in
This section provides a continental synthesis of changes in CIDs, precipitation. Over West Africa, rainfall is projected to decrease
some examples of which are presented in Figure TS.25. in the western Sahel subregion and increase along the Guinea Coast
subregion (medium confidence). Rainfall is projected to increase over
With 2°C global warming, and as early as the mid-21st Eastern Africa (medium confidence). {8.3.1.6, 11.4, 11.9, 12.4.1.2, TS
century, a wide range of CIDs, particularly related to the Atlas.4.2, Atlas.4.4, Atlas.4.5}
water cycle and storms, are expected to show simultaneous
region-specific changes relative to the recent past with high Precipitation declines and aridity trends in Western Africa, Central
or medium confidence. In a number of regions (Southern Africa, Southern Africa and the Mediterranean co-occur with
Africa, the Mediterranean, North Central America, Western trends towards increased agricultural and ecological droughts in
North America, the Amazon regions, South-Western South the same regions (medium confidence). Trends towards increased
America, and Australia), increases in one or more of drought, hydrological droughts have been observed in the Mediterranean
aridity and fire weather (high confidence) will affect a wide (high confidence) and Western Africa (medium confidence). These
range of sectors, including agriculture, forestry, health and trends correspond with projected regional increases in aridity and
ecosystems. In another group of regions (North-Western, fire weather conditions (high confidence). {8.3.1.6, 8.4.1.6, 11.6,
Central and Eastern North America, Arctic regions, North- 11.9, 12.4.1.2}
Western South America, Northern, Western and Central
and Eastern Europe, Siberia, Central, South and East Asia, Wind: Mean wind, extreme winds and the wind energy potential in
Southern Australia and New Zealand), decreases in snow North Africa and the Mediterranean are projected to decrease across
and/or ice or increases in pluvial/river flooding (high all scenarios (high confidence). Over Western Africa and Southern
confidence) will affect sectors such as winter tourism, Africa, a future significant increase in wind speed and wind energy
energy production, river transportation and infrastructure. potential is projected (medium confidence). There is a projected
{11.9, 12.3, 12.4, 12.5, Table 12.2} decrease in the frequency of tropical cyclones making landfall
over Madagascar, East Southern Africa and East Africa (medium
TS.4.3.2.1 Africa confidence). {12.4.1.3}
Additional regional changes in Africa, besides those Snow and ice: There is high confidence that African glaciers and
described in Section TS.4.3.1, include a projected decrease snow have very significantly decreased in the last decades and that
in total precipitation in the northernmost and southernmost this trend will continue in the 21st century. {12.4.1.4}
regions (high confidence), with Western Africa having a west-
to-east pattern of decreasing-to-increasing precipitation Coastal and oceanic: Relative sea level has increased at a higher
(medium confidence). Increases in heavy precipitation that rate than GMSL around Africa over the last 3 decades. The present
can lead to pluvial floods (high confidence) are projected day 1-in-100-year extreme total water level (ETWL) is between
for most African regions, even as increasing dry CIDs 0.1 m and 1.2 m around Africa, with values around 1 m or above
(aridity; hydrological, agricultural and ecological droughts; along the East and West Southern and Central Eastern Africa coasts.
fire weather) are projected in the western part of Western Satellite-derived shoreline retreat rates up to 1 m yr –1 have been
Africa, Southern Africa and Northern Africa and the observed around the continent from 1984 to 2015, except in South
Mediterranean regions (medium to high confidence). {8.4, Eastern Africa, which has experienced a shoreline progradation
11.3, 11.6, 11.9, 12.4, Atlas.4} (growth) rate of 0.1 m yr –1 over the same period. {12.4.1.5}
135
Technical Summary
200
100 NSA
NWS
0 NES
SAM
r.past
r.past
r.past
long.
long.
long.
mid.
mid.
mid.
1.5
1.5
1.5
2
4
2
4
2
4
2.5
NWS NSA NES SWS
0.8
2 0.6 SSA
Recent past value (m3 s-1 km-2)
1.5 0.4
0.2
Change from recent past (m3 s-1 km-2)
1
0.5 0
0.2
0
SWS SAM SES SSA
2.5
0.8
2 0.6
1.5 0.4
1 0.2
0.5 0
0.2
0
1.5 2 4 mid. long. 1.5 2 4 mid. long. 1.5 2 4 mid. long. 1.5 2 4 mid. long.
r.past.
r.past.
r.past.
r.past.
GWL Time slices GWL Time slices GWL Time slices GWL Time slices
136
Technical Summary
(b) 100-yr return period stream flow Maximum temperature exceeding 35°C
NEU WCE MED EEU WSB RAR
0.8 0.1
300 300
days yr -1
0.6
0
0.4 100 100
0.2 0.1 0 0
ESB ECA TIB
0 1.5 2 4 mid. long. 1.5 2 4 mid. long. 1.5 2 4 mid. long. 300 300
days yr -1
r.past.
r.past.
r.past.
GWL Time slices GWL Time slices GWL Time slices 200
200
100 100
0 0
RAR
r.past
r.past
r.past
long.
long.
long.
mid.
mid.
mid.
1.5
1.5
1.5
2
4
2
4
2
4
NEU
EEU WSB ESB RFE GWL Time slices GWL Time slices GWL Time slices
WCE
ECA Shoreline position change
MED WCA
TIB EAS EAS RFE
0.4 200
Change from recent past (m3 s-1 km-2)
1 0.2 100
0 0
0.5 SEA
SAS
–0.2
0 300
Recent past value (m3 s-1 km-2)
days yr -1
1.5
CAF NEAF SEAF 200
0.4 100
1 0.2 0
r.past
r.past
long.
long.
mid.
mid.
0
1.5
1.5
0.5
2
4
2
4
2
Average shoreline
1 0.6
0.2 1.5 200
0.4
0 0
0.5 1 0.2
0 –200
–0.2 0.5
0 –0.2 –400
1.5 2 4 mid. long. 1.5 2 4 mid. long. 0
1.5 2 4 mid. long. –600
r.past.
r.past.
GWL Time slices GWL Time slices mid- long- mid- long- mid- long- mid- long- mid- long-
r.past.
GWL Time slices term term term term term term term term term term
Figure TS.25 | Distribution of projected changes in selected climatic impact-driver (CID) indices for selected regions for Coupled Model Intercomparison
Project Phases 5 and 6 (CMIP6, CMIP5) and Coordinated Regional Downscaling Experiment (CORDEX) model ensembles. The intent of this figure is to show
that many CID projections for multiple global warming levels and scenarios time slices are available for all the AR6 WGI reference regions and are based on both global (CMIP5,
CMIP6) and regional (CORDEX) model ensembles. Different indices are shown for different region: for Eastern Europe and North Asia, the mean number of days per year with
maximum temperature exceeding 35°C; for Central America, the Caribbean, South West Asia, South Asia and South East Asia, the mean number of days per year with the
National Oceanic and Atmospheric Administration (NOAA) Heat Index exceeding 41°C; for Australasia, East Asia and Russian Far East, the average shoreline position change;
for South America, Europe and Africa, the mean change in 1-in-100-year river discharge per unit catchment area (m3 s–1 km–2); and for North America, the median change in
the number of days with snow water equivalent (SWE) over 100 mm. For each box plot, the changes or the climatological values are reported with respect to, or compared to,
the recent past (1995–2014) period for 1.5°C, 2°C and 4°C global warming levels and for mid-century (2041–2060) or end-century (2081–2100) periods for the CMIP5 and
CORDEX RCP8.5 and RCP2.6 and CMIP6 SSP5-8.5 and SSP1-2.6 scenarios ensembles. {Figures 12.5, 12.6, 12.9, 12.SM.1, 12.SM.2, and 12.SM.6}
137
Technical Summary
TS.4.3.2.2 Asia {2.3, 8.3, 8.4, 9.5, 9.6, 10.6, Box 10.4, 11.4, 11.5, 11.7, 11.9,
12.4.2, Atlas.3.1, Atlas.5, Atlas.5.2, Atlas.5.3, Atlas.5.4,
Due to the high climatological and geographical heterogeneity of Atlas.5.5}
Asia, some assessment findings below are summarized over five
sub-continental areas comprising one or more of the AR6 WGI In addition to the main changes summarized above and in Section
reference regions (Box TS.12): East Asia (EAS+ECA), North Asia TS.4.3.1, further details are given below.
(WSB+ESB+RFE), South Asia (SAS), South East Asia (SEA) and South
West Asia (ARP+WCA). Heat and cold: Over all regions of Asia, observed and projected
increases in mean temperature and a shift toward heat extreme
Additional regional changes in Asia, besides those features characteristics are broadly similar to the generic pattern described
described in Section TS.4.3.1, include historical trends in Section TS.4.3.1. Over South East Asia, annual mean surface
of annual precipitation that show considerable regional temperature will likely increase by a slightly smaller amount than the
differences (high confidence). East Asian Monsoon global average. {Atlas.5.4.4}
precipitation has changed, with drying in the north and
wetting in the south since the 1950s, and annual mean Wet and dry: Over East Asia, historical trends of annual precipitation
precipitation totals very likely have increased over show considerable regional differences but with increases over north-
most territories of North Asia since the mid-1970s (high west China and South Korea (high confidence). Daily precipitation
confidence). South Asian summer monsoon precipitation extremes have increased over part of the region (high confidence).
decreased over several areas since the mid-20th century Extreme hydrological drought frequency has increased in a region
(high confidence) but is likely to increase during the 21st extending from south-west to north-east China, with projected
century, with enhanced interannual variability. (Box TS.13) increases of agricultural and ecological drought for 4°C GWL and
fire weather for 2°C and above (medium confidence). {8.3.2, 8.4.2,
TS Increases in precipitation and river floods are projected 11.4.4, 11.4.5, 11.9, 12.4.2.2, Atlas.5.1.2}
over much of Asia: in the annual mean precipitation in
East, North, South and South East Asia (high confidence); Over North Asia, annual mean precipitation totals have very likely
for extremes in East, South, West Central, North and South increased, causing more intense flooding events, and there is
East Asia (high confidence) and Arabian Peninsula (medium medium confidence that the number of dry days has decreased.
confidence); and for river floods in East, South and South Concurrently, total soil moisture is projected to decline extensively
East Asia and East Siberia (medium confidence). Aridity (medium confidence). {8.3.1.3, 8.4.1.6, 11.4.5, 11.5.2, 11.5.5,
in East and West Central Asia is projected to increase, 12.4.2.2, Atlas.5.2.2}
especially beyond the middle of the 21st century and global
warming levels beyond 2°C (medium confidence). Fire Over South Asia, the summer monsoon precipitation decreased
weather seasons are projected to lengthen and intensify over several areas since the mid-20th century (high confidence),
everywhere except South East Asia, Tibetan Plateau and while it increased in parts of the western HKH and decreased over
Arabian Peninsula (medium confidence). eastern-central HKH (medium confidence). The frequency of heavy
precipitation and flood events has increased over several areas
Surface wind speeds have been decreasing in Asia (high during the last few decades (medium confidence). {8.3.1.3, 8.3.2.4.1,
confidence), but there is a large uncertainty in future 8.4.1.5, 8.4.2.4.1, 10.6.3.3, 10.6.3.5, 10.6.3.6, 10.6.3.8, Cross-
trends, with medium confidence that mean wind speeds Chapter Box 10.4, 11.4.1, 11.4.2, 11.4.5, 11.5.5, 12.4.2.2, Box 10.4,
will decrease in North Asia, East Asia and Tibetan Plateau Atlas 5.3.2}
and that tropical cyclones will have decreasing frequency
and increasing intensity overall in South East and East Over South East Asia, mean precipitation trends are not spatially
Asia. coherent or consistent across datasets and seasons (high confidence).
Most of the region has experienced an increase in rainfall intensity
Over North Asia, increases in permafrost temperature and but with a reduced number of wet days (medium confidence). Rainfall
its thawing have been observed over recent decades (high is projected to increase in the northern parts of South East Asia and
confidence). Future projections indicate continuing decline decrease in areas in the Maritime Continent (medium confidence).
in seasonal snow duration, glacial mass, and permafrost area {8.4.1, 11.4.2, 11.5.5, 11.9, 12.4.2.2, Atlas.3.1, Atlas.5.4.2, Atlas.5.4.4}
by mid-century (high confidence). Snow-covered areas and
snow volumes will decrease in most regions of the Hindu Over South West Asia, an observed annual precipitation decline
Kush Himalaya (HKH) during the 21st century, and snowline over the Arabian Peninsula since the 1980s of 6.3 mm per decade
elevations will rise (high confidence) and glacier volumes is contrasted with observed increases between 1.3 mm and 4.8 mm
are likely to decline with greater mass loss in higher CO2 per decade during 1960–2013 over the elevated part of eastern
emissions scenarios. Heavy snowfall is increasing in East West Central Asia (very high confidence), along with an increase of
Asia and North Asia (medium confidence) but with limited the frequency and intensity of extreme precipitation. {Figure 8.19,
evidence on future changes in hail and snow avalanches. Figure 8.20, 8.3.1.6, 8.4.1.6, 11.9, Table 11.2A, 12.4.2.2, Atlas.5.5}
138
Technical Summary
Wind: Over East Asia, the terrestrial near-surface wind speed has TS.4.3.2.3 Australasia
decreased and is projected to decrease further in the future (medium
confidence). Since the mid 1980’s, there has been an increase in the Additional regional changes in Australasia, besides those
number and intensification rate of intense TCs (medium confidence), features described in Section TS.4.3.1, include a significant
with a significant north-westward shift in tracks and a northward decrease in April to October rainfall in the south-west of
shift in their average latitude, increasing exposure over East China, the state of Western Australia, observed from 1910 to 2019
the Korean Peninsula and the Japanese Archipelago (medium and attributable to human influence (high confidence),
confidence). {11.7.1, 12.4.2.3} which is very likely to continue in future. Agricultural
and ecological droughts and hydrological droughts have
Over North Asia, there is medium confidence for a decreasing trend increased over Southern Australia (medium confidence),
in wind speed during 1979–2018 and for projected continuing and meteorological droughts have decreased over Northern
decreases of terrestrial near-surface wind speed. {2.3.1.4.4, 12.4.2.3} and Central Australia (medium confidence). Relative sea
level has increased over the period 1993–2018 at a rate
Over South East Asia, although there is no significant long-term trend higher than GMSL around Australasia (high confidence).
in the number of TCs, fewer but more extreme TCs have affected the Sandy shorelines have retreated around the region, except
Philippines during 1951–2013. {11.7.4, 12.4.2.3} in Southern Australia, where a shoreline progradation rate
of 0.1 m yr –1 has been observed.
Snow and ice: Over East Asia, decreases have been observed in the
frequency, and increases in the mean intensity, of snowfall in north- In the future, heavy precipitation and pluvial flooding are
western, north-eastern and south-eastern China and the eastern very likely to increase over Northern Australia and Central
Tibetan Plateau since the 1960s. Heavy snowfall is projected to Australia, and they are likely to increase elsewhere in
occur more frequently in some parts of Japan (medium confidence). Australasia for global warming levels (GWLs) exceeding 2°C
{12.4.2.4, Atlas.5.1.2} and with medium confidence for a 2°C GWL. Agricultural and TS
ecological droughts are projected to increase in Southern
Over North Asia, seasonal snow duration and extent have decreased and Eastern Australia (medium confidence) for a 2°C GWL.
in recent decades (high confidence), and maximum snow depth likely Fire weather is projected to increase throughout Australia
has increased since the mid-1970s, particularly over the south of the (high confidence) and New Zealand (medium confidence).
Russian Far East. {2.3.2.5, 8.3.1.7.2, 9.5, 12.4.2.4, Atlas.5.2, Atlas.5.4} Snowfall is expected to decrease throughout the region
at high altitudes in both Australia (high confidence) and
Over South Asia, snow cover has reduced over most of the HKH since New Zealand (medium confidence), with glaciers receding
the early 21st century, and glaciers have thinned, retreated, and lost in New Zealand (high confidence). {11.4, Table 11.6, 12.3,
mass since the 1970s (high confidence), although the Karakoram 12.4.3, Atlas.6.4, Atlas.6.5}
glaciers have either slightly gained mass or are in an approximately
balanced state (medium confidence). {8.3.1.7.1, Cross-Chapter In addition to the main changes summarized above and in Section
Box 10.4} TS.4.3.1, further details are given below.
Over South West Asia, mountain permafrost degradation at high Heat and cold: Observed and projected increases in mean
altitudes has increased the instability of mountain slopes in the past temperature and a shift toward heat extreme characteristics are
decade (medium confidence). More than 60% of glacier mass in the broadly similar to the generic pattern described in Section TS.4.3.1.
Caucasus is projected to disappear under RCP8.5 emissions by the {11.9, 12.4.3.1, Atlas.6}
end of the 21st century (medium confidence). {9.5.1, 9.5.3, 12.4.2.4}
Wet and dry: There is medium confidence that heavy precipitation has
Coastal and oceanic: Over the last three decades, relative sea increased in Northern Australia since 1950. Annual mean precipitation
level has increased at a rate higher than GMSL around Asia (high is projected to increase in the south and west of New Zealand (medium
confidence). Gross coastal area loss and shoreline retreat has been confidence) and is projected to decrease in south-west Southern
observed over 1984–2015, but with localized shoreline progradation Australia (high confidence), Eastern Australia (medium confidence), and
in the Russian Far East, East and South East Asia. {12.4.2.5} in the north and east of New Zealand (medium confidence) for a GWL
of 2°C. There is medium confidence that river flooding will increase
Projections show that regional mean sea level continues to rise (high in New Zealand and Australia, with higher increases in Northern
confidence), ranging from 0.4–0.5 m under SSP1-2.6 to 0.8–1.0 m Australia. Aridity is projected to increase with medium confidence in
under SSP5-8.5 for 2081–2100 relative to 1995–2014 (median Southern Australia (high confidence in south-west Southern Australia),
values). This will contribute to more frequent coastal flooding and Eastern Australia (medium confidence) and in the north and east of
higher ETWL in low-lying areas and coastal erosion along sandy New Zealand (medium confidence) for GWLs around 2°C. {11.4, 11.9,
beaches (high confidence). There is high confidence that compound Table 11.6, 12.4.3.2, Atlas.6.2}
effects of climate change, land subsidence, and human factors will
lead to higher flood levels and prolonged inundation in the Mekong Wind: Mean wind speeds are projected to increase in parts of north-
Delta and other Asian coasts. {9.6.1, 9.6.3, 12.4.2.5} eastern Australia (medium confidence) by the end of the 21st century
139
Technical Summary
under high CO2 emissions scenarios. TCs in north-eastern and north South America and Southern Central America (medium confidence).
Australia are projected to decrease in number (high confidence) but In Northern South America and Southern Central America, aridity
increase in intensity except for ‘east coast lows’ (low confidence). and agricultural and ecological droughts are increasing with
{12.4.3.3} medium confidence. Fire weather is projected to increase over
Southern Central America and Southern South America with medium
Snow and ice: Observations in Australia show that the snow season confidence. {8.3.1.3, 8.4.2.4.5, 11.4.2, 11.9, Table 11.14, Table 11.15,
length has decreased by 5% in the last five decades. Furthermore, the 12.4.4.2, Atlas.7.2.2, Atlas.7.2.4}
date of peak snowfall in Australia has advanced by 11 days over the
last 5 decades. Glacier ice volume in New Zealand has decreased by Wind: Climate projections indicate an increase in mean wind speed
33% from 1977 to 2018. {12.4.3.4, Atlas.6.2} and in wind power potential over the Amazonian region (Northern
South America, South American Monsoon, North-Eastern South
Coastal and oceanic: Observed changes in marine heatwaves America) (medium confidence). {12.4.4.3}
(MHWs) over the 20th century in the region show an increase in their
occurrence frequency, except along the south-east coast of New Snow and ice: Glacier volume loss and permafrost thawing will
Zealand, an increase in duration per event, and the total number of likely continue in the Andes Cordillera under all climate scenarios,
MHW days per decade, with the change being stronger in the Tasman causing important reductions in river flow and potentially high-
Sea than elsewhere. The present day 1-in-100-year ETWL is between magnitude glacial lake outburst floods. {9.5.1.1, 12.4.4.4}
0.5–2.5 m around most of Australia, except the north-western
coast where 1-in-100-year ETWL can be as high as 6–7 m. {Box 9.1, Coastal and oceanic: Around Central and South America, relative
12.3.1.5, 12.4.3.5} sea level has increased at a higher rate than GMSL in the South
Atlantic and the subtropical North Atlantic, and at a rate lower than
TS.4.3.2.4 Central and South America GMSL in the East Pacific over the last 3 decades. The present day
TS 1-in-100-year ETWL is highest in Southern and South-Western South
Additional regional changes in Central and South America, America subregions, where it can be as large as 5 to 6 m. Satellite
besides those features described in Section TS.4.3.1, include observations for 1984–2015 show shoreline retreat rates along the
increases in mean and extreme precipitation in South- sandy coasts of Southern Central America, South-Eastern South
Eastern South America since the 1960s (high confidence) America and Southern South America, while shoreline progradation
(Section TS.4.2.3). Decreasing trends in mean precipitation rates have been observed in North-Western South America and
and increasing trends in agricultural and ecological drought Northern South America. Over the period 1982–2016, the coastlines
are observed over North-Eastern South America (medium experienced at least one MHW per year, and more along the Pacific
confidence). The intensity and frequency of extreme coast of North Central America and the Atlantic coast of South-
precipitation and pluvial floods is projected to increase Eastern South America. {12.4.4.5}
over South-Eastern South America, Southern South America,
Northern South America, South American Monsoon and TS.4.3.2.5 Europe
North-Eastern South America (medium confidence) for a 2°C
GWL and above. Increases of agricultural and ecological Additional regional changes in Europe, besides those features
drought are projected in South America Monsoon and described in Section TS.4.3.1, include observed increases
Southern South America, and fire weather is projected to in pluvial flooding in Northern Europe and hydrological
increase over several regions (Northern South America, the and agricultural/ecological droughts in the Mediterranean
South American Monsoon, North-Eastern South America (high confidence), which have been attributed to human
and South-Western South America) (high confidence). {8.3, influence with high and medium confidence, respectively.
8.4, 11.3, 11.4, 11.9, Table 11.13, Table 11.14, Table 11.15, Increased mean precipitation amounts at high latitudes
12.4.4.2, Atlas.7.1, Atlas.7.2} in boreal winter and reduced summer precipitation in
southern Europe are projected starting from a 2°C GWL
In addition to the main changes summarized above and in Section (high confidence). Aridity, agricultural and hydrological
TS.4.3.1, further details are given below. droughts and fire weather conditions will increase in
the Mediterranean region starting from 2°C GWL (high
Heat and cold: Observed and projected increases in mean confidence). Pluvial flooding will increase everywhere
temperature and a shift toward heat extreme characteristics are with high confidence except for medium confidence in the
broadly similar to the generic pattern described in Section TS.4.3.1. Mediterranean; in Western and Central Europe this also
{11.3.2, 11.3.5, Table 11.13, 12.4.4.1, Atlas.7.1.2, Atlas.7.2.2, applies to river flooding starting from a 2°C GWL (high
Atlas.7.2.4} confidence). Most periglacial processes in Northern Europe
are projected to disappear by the end of the 21st century,
Wet and dry: Mean precipitation is projected to change in a dipole even for a low warming scenario (medium confidence). {8.3,
pattern with increases in North-Western and South-Eastern South 11.3, 11.9, 12.4.5, 12.5.2, Atlas.8.2, Atlas.8.4}
America and decreases in North-Eastern and South-Western South
America (high confidence) and with further decreases in Northern
140
Technical Summary
In addition to the main changes summarized above and in Section America and Northern Central America (from medium to
TS.4.3.1, further details are given below. high confidence). Severe wind storms, tropical cyclones
and dust storms in North America are shifting toward more
Heat and cold: Observed and projected increases in mean extreme characteristics (medium confidence), and both
temperature and a shift toward heat extreme characteristics are observations and projections point to strong changes in the
broadly similar to the generic pattern described in Section TS.4.3.1. seasonal and geographic range of snow and ice conditions
{11.3, 11.9, 12.4.5.1, 12.5.2, Atlas.8.2, Atlas.8.4} in the coming decades (very high confidence). General
findings for relative sea level, coastal flooding and erosion
Wet and dry: There is medium confidence that annual mean will not apply for areas with substantial land uplift around
precipitation has increased in Northern Europe, West and Central the Hudson Bay and Southern Alaska. {8.4, 11.4, 11.5, 11.7,
Europe, and Eastern Europe since the early 20th century and 11.9, 12.4, Atlas.9.4}
high confidence for increases in extreme precipitation. In the
European Mediterranean, the magnitude and sign of observed land In addition to the main changes summarized above and in Section
precipitation trends depend on time period and exact study region TS.4.3.1, further details are given below.
(medium confidence). There is medium confidence that river floods
will decrease in Northern, Eastern and southern Europe for high Heat and cold: Observed and projected increases in mean
warming levels. {8.3.1.3, 11.3, 11.9, 12.4.5.2, Atlas.8.2, Atlas.8.4} temperature and a shift toward heat extreme characteristics are
broadly similar to the generic pattern described in Section TS.4.3.1.
Wind: Mean wind speed over land has decreased (medium {11.3, 11.9, 12.4.6.1, Atlas.9.2, Atlas.9.4}
confidence), but the role of human-induced climate change has not
been established. There is high confidence that mean wind speeds Wet and dry: Annual precipitation increased over parts of Eastern
will decrease in Mediterranean areas and medium confidence for and Central North America during 1960–2015 (high confidence) and
such decreases in Northern Europe for GWLs exceeding 2°C. The has decreased in parts of south-western United States and north- TS
frequency of Medicanes (tropical-like cyclones in the Mediterranean) western Mexico (medium confidence). River floods are projected to
is projected to decrease (medium confidence). {11.9, 12.4.5.3} increase for all North American regions other than Northern Central
America (medium confidence). {8.4.2.4, 11.4, 11.5, 11.9, 12.4.6.2,
Snow and ice: In the Alps, snow cover will decrease below elevations Atlas.9.2, Atlas.9.4}
of 1500–2000 m throughout the 21st century (high confidence).
A reduction of glacier ice volume is projected in the European Alps Agricultural and ecological drought increases have been observed in
and Scandinavia with high confidence and with medium confidence Western North America (medium confidence), and aridity is projected
for the timing and mass change rates. {9.5.2, 12.4.5.4} to increase in the south-western United States and Northern Central
America, with lower summer soil moisture across much of the
Coastal and oceanic: Over the last three decades, relative sea level continental interior (medium confidence). {8.4.1, 11.6.2, 12.4.6.2}
has increased at a lower rate than GMSL in the sub-polar North
Atlantic coasts of Europe. The present-day 1-in-100-year ETWL is Wind: Projections indicate a greater number of the most intense
between 0.5–1.5 m in the Mediterranean basin and 2.5–5.0 m in TCs, with slower translation speeds and higher rainfall potential for
the western Atlantic European coasts, around the United Kingdom Mexico’s Pacific Coast, the Gulf Coast and the United States East
and along the North Sea coast, and lower at 1.5–2.5 m along the Coast (medium confidence). Mean wind speed and wind power
Baltic Sea coast. Satellite-derived shoreline change estimates over potential are projected to decrease in Western North America (high
1984–2015 indicate shoreline retreat rates of around 0.5 m yr –1 confidence), with differences between global and regional models
along the sandy coasts of Central Europe and the Mediterranean lending low confidence elsewhere. {11.4, 11.7, 12.4.6.3}
and more or less stable shorelines in Northern Europe. Over the
period 1982–2016, the coastlines of Europe experienced on average Snow and ice: It is likely that some high-latitude regions will
more than 2.0 MHW per year, with the eastern Mediterranean and experience an increase in winter snow water equivalent due to the
Scandinavia experiencing 2.5–3 MHWs per year. {12.4.5.5} snowfall increase prevailing over the warming trend. At sustained
GWLs between 3°C and 5°C, nearly all glacial mass in Western
TS.4.3.2.6 North America Canada and Western North America will disappear (medium
confidence). {9.5.1, 9.5.3, 12.4.6.4, Atlas.9.4}
Additional regional changes in North America, besides those
features described in Section TS.4.3.1, include changes Coastal and oceanic: Around North America, relative sea level has
in North American wet and dry CIDs, which are largely increased over the last three decades at a rate lower than GMSL
organized by the north-east (more wet) to south-west (more in the subpolar North Atlantic and in the East Pacific, while it has
dry) pattern of mean precipitation change, although heavy increased at a rate higher than GMSL in the subtropical North Atlantic.
precipitation increases are widespread (high confidence). Observations indicate that episodic coastal flooding is increasing
Increasing evaporative demand will expand agricultural along many coastlines in North America. Shoreline retreat rates of
and ecological drought and fire weather (particularly in around 1 m yr –1 have been observed during 1984–2015 along the
summertime) in Central North America, Western North sandy coasts of North-Western North America and Northern Central
141
Technical Summary
America, while portions of the United States Gulf Coast have seen Permafrost warming, loss of seasonal snow cover, and
a retreat rate approaching 2.5 m yr –1. Sandy shorelines along Eastern glacier melt will be widespread (high confidence). There is
North America and Western North America have remained more or high confidence that both the Greenland and Antarctic ice
less stable during 1984–2014, but a shoreline progradation rate of sheets have lost mass since 1992 and will continue to lose
around 0.5 m yr –1 has been observed in North-Eastern North America. mass throughout this century under all emissions scenarios.
{12.4.6.5} Relative sea level and coastal flooding are projected to
increase in areas other than regions with substantial land
TS.4.3.2.7 Small Islands uplift (medium confidence). {2.3, 3.4, 4.3, 4.5, 7.4, 8.2, 8.4,
Box 8.2, 9.5, 12.4.9, Atlas.11.1, Atlas.11.2}
Additional regional changes in Small Islands, besides those
features described in Section TS.4.3.1, include a likely In addition to the main changes summarized above and in Section
decrease in rainfall during boreal summer in the Caribbean TS.4.3.1, further details are given below.
and in some parts of the Pacific islands poleward of 20°
latitude in both the Northern and Southern Hemispheres. Heat and cold: Changes in Antarctica showed larger spatial
These drying trends will likely continue in coming decades. variability, with very likely warming in the Antarctic Peninsula since
Fewer but more intense tropical cyclones are projected the 1950s and no overall trend in East Antarctica. Less warming
starting from a 2°C GWL (medium confidence). {9.6, 11.3, and weaker polar amplification are projected as very likely over
11.4, 11.7, 11.9, 12.4.7, Atlas.10.2, Atlas.10.4, Cross-Chapter the Antarctic than in the Arctic, with a weak polar amplification
Box Atlas.2} projected as very likely by the end of the 21st century. {4.3.1, 4.5.1,
7.4.4, 12.4.9.1, Atlas.11.1, Atlas.11.2}
In addition to the main changes summarized above and in Section
TS.4.3.1, further details are given below. Wet and dry: Recent decades have seen a general decrease in
TS Arctic aridity (high confidence), with increased moisture transport
Heat and cold: It is very likely that most Small Islands have warmed leading to higher precipitation, humidity and streamflow and a
over the period of instrumental records, and continued temperature corresponding decrease in dry days. Antarctic precipitation showed
increases in the 21st century will further increase heat stress in these a positive trend during the 20th century. The water cycle is projected
regions. {11.3.2, 11.9, 12.4.7.1, Atlas.10.2, Atlas.10.4, Cross-Chapter to intensify in both polar regions, leading to higher precipitation
Box Atlas.2} totals (and a shift to more heavy precipitation) and higher fraction
of precipitation falling as rain. In the Arctic, this will result in higher
Wet and dry: Observed and projected rainfall trends vary spatially river flood potential and earlier meltwater flooding, altering seasonal
across the Small Islands. Higher evapotranspiration under a warming characteristics of flooding (high confidence). A lengthening of the fire
climate can partially offset future increases or amplify future season (medium confidence) and encroachment of fire regimes into
reductions in rainfall, resulting in increased aridity as well as more tundra regions (high confidence) are projected. {8.2.3, 8.4.1, Box 8.2,
severe agricultural and ecological drought in the Caribbean (medium 9.4.1, 9.4.2, 12.4.9.2, Atlas.11.1, Atlas.11.2}
confidence). {11.4.2, 11.9, 12.4.7.2, Atlas.10.2, Atlas.10.4, Cross-
Chapter Box Atlas.2} Wind: There is medium confidence in mean wind decrease over
the Russian Arctic and Arctic North-East North America, but low
Wind: Global changes indicate that Small Islands will face fewer confidence of changes in other Arctic regions and Antarctica.
but more intense TCs, with spatial inconsistency in projections given {12.4.9.3}
poleward shifts in TC tracks (medium confidence). {11.7.1.2, 11.7.1.5,
12.4.7.3} Snow and ice: Reductions in spring snow cover extent have
occurred across the Northern Hemisphere since at least 1978 (very
Coastal and oceanic: Continued relative sea level rise is very likely high confidence). Permafrost warming and thawing have been
in the ocean around Small Islands and, along with storm surges widespread in the Arctic since the 1980s (high confidence), causing
and waves, will exacerbate coastal inundation with the potential to strong heterogeneity in surface conditions. There is high confidence
increase saltwater intrusion into aquifers in small islands. Shoreline in future glacier- and ice-sheet loss, permafrost warming, decreasing
retreat is projected along sandy coasts of most small islands (high permafrost extent and decreasing seasonal duration and extent of
confidence). {9.6.3.3, 12.4.7.4, Cross-Chapter Box Atlas.2} snow cover in the Arctic. Decline in seasonal sea ice coverage along
the majority of the Arctic coastline in recent decades is projected to
TS.4.3.2.8 Polar continue, contributing to an increase in coastal hazards (including
open water storm surge, coastal erosion and flooding). {2.3.2, 3.4.2,
It is virtually certain that surface warming in the Arctic 3.4.3, 9.4.1, 9.4.2, 9.5, 12.4.6, 12.4.9, Atlas.11.2}
will continue to be more pronounced than the global
average warming over the 21st century. An intensification Coastal and oceanic: Higher sea levels contribute to high
of the polar water cycle will increase mean precipitation, confidence for projected increases of Arctic coastal flooding and
with precipitation intensity becoming stronger and more higher coastal erosion (aided by sea ice loss) (medium confidence),
likely to be rainfall rather than snowfall (high confidence).
142
Technical Summary
with lower confidence for those regions with substantial land uplift TS.4.3.2.10 Other Typological Domains
(Arctic North-East North America and Greenland). {12.4.9.5}
Some types of regions found in different continents face
TS.4.3.2.9 Ocean common climate challenges regardless of their location.
These include biodiversity hot spots that will very likely
The Indian Ocean, western equatorial Pacific Ocean and see even more extreme heat and droughts, mountain areas
western boundary currents have warmed faster than the where a projected raising in the freezing level height will
global average (very high confidence), with the largest alter snow and ice conditions (high confidence), and tropical
changes in the frequency of marine heatwaves (MHWs) forests that are increasingly prone to fire weather (medium
projected in the western tropical Pacific and the Arctic confidence). {8.4, Box 8.2, 9.5, 12.3, 12.4}
Ocean (medium confidence). The Pacific and Southern Ocean
are projected to freshen and the Atlantic to become more Biodiversity hotspots located around the world will each face unique
saline (medium confidence). Anthropogenic warming is very challenges in CID changes. Heat, drought and length of dry season,
likely to further decrease ocean oxygen concentrations, and wildfire weather, sea surface temperature and deoxygenation are
this deoxygenation is expected to persist for thousands relevant drivers to terrestrial and freshwater ecosystems and have
of years (medium confidence). Arctic sea ice losses are marked increasing trends. {12.3, 12.4.10.1}
projected to continue, leading to a practically ice-free Arctic
in September by the end of the 21st century under high Desert and semi-arid areas are strongly affected by CIDs such as
CO2 emissions scenarios (high confidence). {2.3, 5.3, 9.2, 9.3, extreme heat, drought and dust storms, with large-scale aridity
Box 9.2, 12.3.6, 12.4.8} trends contributing to expanding drylands in some regions (high
confidence). {12.3, 12.4.10.3}
In addition to the main changes summarized above and in Section
TS.4.3.1, further details are given below. Average warming in mountain areas varies with elevation, but TS
the pattern is not globally uniform (medium confidence). Extreme
Ocean surface temperature: The Southern Ocean, the eastern precipitation is projected to increase in major mountainous regions
equatorial Pacific, and the North Atlantic Ocean have warmed more (medium to high confidence depending on location), with potential
slowly than the global average or slightly cooled. Global warming cascading consequences of floods, landslides and lake outbursts
of 2°C above 1850–1900 levels would result in the exceedance of in all scenarios (medium confidence). {8.4.1.5, Box 8.2, 9.5.1.3,
numerous hazard thresholds for pathogens, seagrasses, mangroves, 9.5.3.3, 9.5.2.3, Cross-Chapter Box 10.4, 11.5.5, 12.3, 12.4.1–12.4.6,
kelp forests, rocky shores, coral reefs and other marine ecosystems 12.4.10.4}
(medium confidence). {9.2.13, 12.4.8}
Most tropical forests are challenged by a mix of emerging warming
Marine heatwaves: Moderate increases in MHW frequency are trends that are particularly large in comparison to historical variability
projected for mid-latitudes, and only small increases are projected (medium confidence). Water cycle changes bring prolonged drought,
for the Southern Ocean (medium confidence). Under the SSP5-8.5 longer dry seasons and increased fire weather to many tropical
scenario, permanent MHWs (more than 360 days per year) are forests (medium confidence). {10.5, 12.3, 12.4}
projected to occur in the 21st century in parts of the tropical ocean,
the Arctic Ocean, and around 45°S; however, the occurrence of
such permanent MHWs can be largely avoided under the SSP1-2.6
scenario. {Box 9.2, 12.4.8}
Ocean acidity: With the rising CO2 concentration, the ocean surface
pH has declined globally over the past four decades (virtually certain).
{2.3.3.5, 5.3.3.2, 12.4.8}
Ocean salinity: At the basin scale, it is very likely that the Pacific and
the Southern Ocean have freshened while the Atlantic has become
more saline. {2.3.3.2, 9.2.2.2, 12.4.8}
Sea ice: Arctic perennial sea ice is being replaced by thin, seasonal
ice, with earlier spring melt and delayed fall freeze up. There is no
clear trend in the Antarctic sea ice area over the past few decades
and low confidence in its future change. {2.3.2.1.1, 9.3.1.1, 12.4.8,
12.4.9}
143
Technical Summary
With global warming, urban areas and cities will be affected by more frequent occurrences of extreme climate events,
such as heatwaves, with more hot days and warm nights as well as sea level rise and increases in tropical cyclone
storm surge and rainfall intensity that will increase the probability of coastal city flooding (high confidence). {Box 10.3,
11.3, 11.5, 12.3, 12.4}
Urban areas have special interactions with the climate system, for instance in terms of heat islands and altering the water cycle, and
thereby will be more affected by extreme climate events such as extreme heat (high confidence). With global warming, increasing
relative sea level compounded by increasing tropical cyclone storm surge and rainfall intensity will increase the probability of coastal
city flooding (high confidence). Arctic coastal settlements are particularly exposed to climate change due to sea ice retreat (high
confidence). Improvements in urban climate modelling and climate monitoring networks have contributed to understanding the
mutual interaction between regional and urban climate (high confidence). {Box 10.3, 11.3, 11.5, 12.3, 12.4}
Despite having a negligible effect on global surface temperature (high confidence), urbanization has exacerbated the effects of global
warming through its contribution to the observed warming trend in and near cities, particularly in annual mean minimum temperature
(very high confidence) and increases in mean and extreme precipitation over and downwind of the city, especially in the afternoon
and early evening (medium confidence). {2.3, Box 10.3, 11.3, 11.4, 12.3, 12.4}
Combining climate change projections with urban growth scenarios, future urbanization will amplify (very high confidence) the
projected local air temperature increase, particularly by strong influence on minimum temperatures, which is approximately comparable
TS in magnitude to global warming (high confidence). Compared to present day, large implications are expected from the combination
of future urban development and more frequent occurrence of extreme climate events, such as heatwaves, with more hot days and
warm nights adding to heat stress in cities (very high confidence). {Box 10.2, 11.3, 12.4}
Both sea levels and air temperatures are projected to rise in most coastal settlements (high confidence). There is high confidence in
an increase in pluvial flood potential in urban areas where extreme precipitation is projected to increase, especially at high global
warming levels. {11.4, 11.5, 12.4}
144
Frequently Asked Questions
FAQ Frequently Asked Questions
Coordinating Editors:
Sophie Berger (France/Belgium), Sarah L. Connors (France/United Kingdom)
Drafting Authors:
Richard P. Allan (United Kingdom), Paola A. Arias (Colombia), Kyle Armour (United States of America),
Terje Berntsen (Norway), Lisa Bock (Germany), Ruth Cerezo-Mota (Mexico), Kim Cobb (United States
of America), Alejandro Di Luca (Australia, Canada/Argentina), Paul Edwards (United States of America),
Tamsin L. Edwards (United Kingdom), Seita Emori (Japan), François Engelbrecht (South Africa), Veronika
Eyring (Germany), Piers Forster (United Kingdom), Baylor Fox- Kemper (United States of America),
Sandro Fuzzi (Italy), John C. Fyfe (Canada), Nathan P. Gillett (Canada), Nicholas R. Golledge (New
Zealand/United Kingdom), Melissa I. Gomis (France/Switzerland), William J. Gutowski (United States
of America), Rafiq Hamdi (Belgium), Mathias Hauser (Switzerland), Ed Hawkins (United Kingdom),
Nigel Hawtin (United Kingdom), Darrell S. Kaufman (United States of America), Megan Kirchmeier-
Young (Canada/ United States of America), Charles Koven (United States of America), June-Yi Lee
(Republic of Korea), Sophie Lewis (Australia), Jochem Marotzke (Germany), Valérie Masson-Delmotte
(France), Thorsten Mauritsen (Sweden/Denmark), Thomas K. Maycock (United States of America),
Shayne McGregor (Australia), Sebastian Milinski (Germany), Olaf Morgenstern (New Zealand/
Germany), Swapna Panickal (India), Joeri Rogelj (United Kingdom/Belgium), Maisa Rojas (Chile), Alex
C. Ruane (United States of America), Bjørn H. Samset (Norway), Trude Storelvmo (Norway), Sophie
Szopa (France), Jessica Tierney (United States of America), Russell S. Vose (United States of America),
Masahiro Watanabe (Japan), Sönke Zaehle (Germany), Xuebin Zhang (Canada), Kirsten Zickfeld
(Canada/Germany)
These Frequently Asked Questions have been extracted from the chapters of the underlying report and are compiled
here. When referencing specific FAQs, please reference the corresponding chapter in the report from where the FAQ
originated (e.g., FAQ 3.1 is part of Chapter 3).
147
Table of Contents
Frequently Asked Questions FAQ 7.1 What Is the Earth’s Energy Budget,
and What Does It Tell Us
FAQ 1.1 Do We Understand About Climate Change? ...................... 184
Climate Change Better Now
Compared to When the IPCC Started? 150 FAQ 7.2 What Is the Role of Clouds
in a Warming Climate? ........................ 186
FAQ 1.2 Where Is Climate Change
Most Apparent? .................................. 152 FAQ 7.3 What Is Equilibrium Climate Sensitivity
and How Does It Relate
FAQ 1.3 What Can Past Climate Teach Us to Future Warming? ............................ 188
About the Future? .............................. 154
FAQ 8.1 How Does Land Use Change
FAQ 2.1 The Earth’s Temperature Alter the Water Cycle? ........................ 190
Has Varied Before. How Is the
Current Warming Any Different? ........ 156 FAQ 8.2 Will Floods Become More Severe
or More Frequent as a Result
FAQ 2.2 What Is the Evidence of Climate Change? ............................ 192
for Climate Change? ........................... 158
FAQ 8.3 What Causes Droughts, and Will Climate
FAQ 3.1 How Do We Know Humans Change Make Them Worse? ............... 194
Are Responsible for Climate Change? 160
FAQ 9.1 Can Continued Melting of the Greenland
FAQ 3.2 What is Natural Variability and Antarctic Ice Sheets Be Reversed?
and How Has it Influenced How Long Would It Take for Them
Recent Climate Changes? ................... 162 to Grow Back? .................................... 196
FAQ 3.3 Are Climate Models Improving? ......... 164 FAQ 9.2 How Much Will Sea Level Rise
in the Next Few Decades? .................. 198
FAQ 4.1 How Will the Climate Change
Over the Next Twenty Years? .............. 166 FAQ 9.3 Will the Gulf Stream Shut Down? ....... 200
FAQ 4.2 How Quickly Would We See FAQ 10.1 How Can We Provide
the Effects of Reducing Useful Climate Information
Carbon Dioxide Emissions? ................ 168 for Regional Stakeholders? ................ 202
FAQ 4.3 At a Given Level of Global Warming, FAQ 10.2 Why Are Cities Hotspots
What Are the Spatial Patterns of Global Warming? ............................ 204
of Climate Change? ............................ 170
FAQ 11.1 How Do Changes In Climate Extremes
FAQ 5.1 Is the Natural Removal of Carbon Compare With Changes
From the Atmosphere Weakening? ..... 172 In Climate Averages? .......................... 206
FAQ 5.2 Can Thawing Permafrost Substantially FAQ 11.2 Will Unprecedented Extremes Occur
Increase Global Warming? .................. 174 As a Result Of Human-Induced
Climate Change? ................................. 208
FAQ 5.3 Could Climate Change Be Reversed
By Removing Carbon Dioxide FAQ 11.3 Did Climate Change Cause
From the Atmosphere? ........................ 176 That Recent Extreme Event
In My Country? .................................... 209
FAQ 5.4 What Are Carbon Budgets? ................ 178
FAQ 12.1 What Is a
FAQ 6.1 What Are Short-lived Climate Forcers Climatic Impact-driver (CID)? ............. 211
and How Do They Affect the Climate? 180
FAQ 12.2 What Are Climatic Thresholds
FAQ 6.2 What Are the Links Between and Why Are They Important? ............ 213
Limiting Climate Change
and Improving Air Quality? ................ 182 FAQ 12.3 How Will Climate Change Affect
the Regional Characteristics
of a Climate Hazard? ........................ 215
149
Frequently Asked Questions
FAQ 1.1 | Do We Understand Climate Change Better Now Compared to When the IPCC Started?
Yes, much better. The first IPCC report, released in 1990, concluded that human-caused climate change would
soon become evident, but could not yet confirm that it was already happening. Today, evidence is overwhelming
that the climate has indeed changed since the pre-industrial era and that human activities are the principal cause
of that change. With much more data and better models, we also understand more about how the atmosphere
interacts with the ocean, ice, snow, ecosystems and land surfaces of the Earth. Computer climate simulations
have also improved dramatically, incorporating many more natural processes and providing projections at much
higher resolutions.
Since the first IPCC report in 1990, large numbers of new instruments have been deployed to collect data in the
air, on land, at sea and from outer space. These instruments measure temperature, clouds, winds, ice, snow,
ocean currents, sea level, soot and dust in the air, and many other aspects of the climate system. New satellite
instruments have also provided a wealth of increasingly fine-grained data. Additional data from older observing
systems and even hand-written historical records are still being incorporated into observational datasets, and
these datasets are now better integrated and adjusted for historical changes in instruments and measurement
techniques. Ice cores, sediments, fossils, and other new evidence from the distant past have taught us much
about how Earth’s climate has changed throughout its history.
Understanding of climate system processes has also improved. For example, in 1990 very little was known about
how the deep ocean responds to climate change. Today, reconstructions of deep-ocean temperatures extend as
far back as 1871. We now know that the oceans absorb most of the excess energy trapped by greenhouse gases
and that even the deep ocean is warming up. As another example, in 1990, relatively little was known about
exactly how or when the gigantic ice sheets of Greenland and Antarctica would respond to warming. Today,
much more data and better models of ice-sheet behaviour reveal unexpectedly high melt rates that will lead to
major changes within this century, including substantial sea level rise (FAQ 9.2).
The major natural factors contributing to climate change on time scales of decades to centuries are volcanic
eruptions and variations in the sun’s energy output. Today, data show that changes in incoming solar energy
since 1900 have contributed only slightly to global warming, and they exhibit a slight downward trend since the
FAQ 1970s. Data also show that major volcanic eruptions have sometimes cooled the entire planet for relatively short
periods of time (typically several years) by erupting aerosols (tiny airborne particles) high into the atmosphere.
The main human causes of climate change are the heat-absorbing greenhouse gases released by fossil fuel
combustion, deforestation, and agriculture, which warm the planet; and aerosols such as sulphate from burning
coal, which have a short-term cooling effect that partially counteracts human-caused warming. Since 1990, we
have more and better observations of these human factors as well as improved historical records, resulting in
more precise estimates of human influence on the climate system (FAQ 3.1).
While most climate models in 1990 focused on the atmosphere, using highly simplified representations of oceans
and land surfaces, today’s Earth system simulations include detailed models of oceans, ice, snow, vegetation and
many other variables. An important test of models is their ability to simulate Earth’s climate over the period
of instrumental records (since about 1850). Several rounds of such testing have taken place since 1990, and
the testing itself has become much more rigorous and extensive. As a group and at large scales, models have
predicted the observed changes well in these tests (FAQ 3.3). Since there is no way to do a controlled laboratory
experiment on the actual Earth, climate model simulations can also provide a kind of ‘alternate Earth’ to test
what would have happened without human influence. Such experiments show that the observed warming
would not have occurred without human influence.
Finally, physical theory predicts that human influence on the climate system should produce specific patterns
of change, and we see those patterns in both observations and climate simulations. For example, nights are
warming faster than days, less heat is escaping to space, and the lower atmosphere (troposphere) is warming but
the upper atmosphere (stratosphere) has cooled. These confirmed predictions are all evidence of changes driven
primarily by increases in GHG concentrations rather than natural causes.
150
Frequently Asked Questions
FAQ 1.1: Do we understand climate change better than when the IPCC started?
Yes. Between 1990 and 2021, observations, models and climate understanding improved, while the dominant
role of human influence in global warming was confirmed.
1990 2021
IPCC IPCC
First Sixth
Assessment Assessment
Understanding
Human influence on climate ? Suspected Established fact
Observations
Global warming since late 1800s 0.3–0.6°C 0.95–1.20°C
Satellite remote sensing Temperature, snow cover, Temperature, cryosphere, Earth radiation budget, CO2,
Earth radiation budget sea level, clouds, aerosols, land cover, many others
FAQ
Climate models Global Global Regional
Major elements Circulating atmosphere and ocean Circulating atmosphere and ocean
Atmospheric chemistry
Land use/cover
FAQ 1.1, Figure 1 | Sample elements of climate understanding, observations and models as assessed in the IPCC First Assessment Report
(1990) and Sixth Assessment Report (2021). Many other advances since 1990, such as key aspects of theoretical understanding, geological records and
attribution of change to human influence, are not included in this figure because they are not readily represented in this simple format. Fuller explanations of
the history of climate knowledge are available in the introductory chapters of the IPCC Fourth and Sixth assessment reports.
151
Frequently Asked Questions
The signs of climate change are unequivocal at the global scale and are increasingly apparent on smaller spatial
scales. The high northern latitudes show the largest temperature increase, with clear effects on sea ice and
glaciers. The warming in the tropical regions is also apparent because the natural year-to-year variations in
temperature there are small. Long-term changes in other variables such as rainfall and some weather and climate
extremes have also now become apparent in many regions.
It was first noticed that the planet’s land areas were warming in the 1930s. Although increasing atmospheric
carbon dioxide (CO2) concentrations were suggested as part of the explanation, it was not certain at the time
whether the observed warming was part of a long-term trend or a natural fluctuation: global warming had not
yet become apparent. But the planet continued to warm, and by the 1980s the changes in temperature had
become obvious or, in other words, the signal had emerged.
Imagine you had been monitoring temperatures at the same location for the past 150 years. What would you
have experienced? When would the warming have become noticeable in your data? The answers to these
questions depend on where on the planet you are.
Observations and climate model simulations both demonstrate that the largest long-term warming trends are in
the high northern latitudes and the smallest warming trends over land are in tropical regions. However, the year-
to-year variations in temperature are smallest in the tropics, meaning that the changes there are also apparent,
relative to the range of past experiences (FAQ 1.2, Figure 1).
Changes in temperature also tend to be more apparent over land areas than over the open ocean and are often
most apparent in regions which are more vulnerable to climate change. It is expected that future changes will
continue to show the largest signals at high northern latitudes, but with the most apparent warming in the
tropics. The tropics also stand to benefit the most from climate change mitigation in this context, as limiting
global warming will also limit how far the climate shifts relative to past experience.
Changes in other climate variables have also become apparent at smaller spatial scales. For example, changes
in average rainfall are becoming clear in some regions, but not in others, mainly because natural year-to-year
FAQ variations in precipitation tend to be large relative to the magnitude of the long-term trends. However, extreme
rainfall is becoming more intense in many regions, potentially increasing the impacts from inland flooding (FAQ
8.2). Sea levels are also clearly rising on many coastlines, increasing the impacts of inundation from coastal storm
surges, even without any increase in the number of storms reaching land. A decline in the amount of Arctic sea
ice is apparent, both in the area covered and in its thickness, with implications for polar ecosystems.
When considering climate-related impacts, it is not necessarily the size of the change that is most important.
Instead, it can be the rate of change or it can also be the size of the change relative to the natural variations
of the climate to which ecosystems and society are adapted. As the climate is pushed further away from past
experiences and enters an unprecedented state, the impacts can become larger, along with the challenge of
adapting to them.
How and when a long-term trend becomes distinguishable from shorter-term natural variations depends on the
aspect of climate being considered (e.g., temperature, rainfall, sea ice or sea level), the region being considered,
the rate of change, and the magnitude and timing of natural variations. When assessing the local impacts from
climate change, both the size of the change and the amplitude of natural variations matter.
152
Frequently Asked Questions
Estimation of:
2 standard deviations of natural year-to-year variations
1 standard deviation of natural year-to-year variations
High latitudes (e.g. mid-North America) Low latitudes (e.g. Tropical South America)
°C
Temperature change since 1850-1900 (°C)
2.0 2.0
1.5 1.5
Smaller trend that
deviates earlier
1.0 1.0 from past conditions
0.5 0.5
0 0
-0.5 -0.5
-1.0 -1.0
1860 1890 1920 1950 1980 2010 1860 1890 1920 1950 1980 2010
FAQ 1.2, Figure 1 | Observed variations in regional temperatures since 1850 (data from Berkeley Earth). Regions in high latitudes, such as
mid-North America (40°N–64°N, 140°W–60°W, left), have warmed by a larger amount than regions at lower latitudes, such as tropical South America
(10°S–10°N, 84°W–16°W, right), but the natural variations are also much larger at high latitudes (darker and lighter shading represents 1 and 2 standard
deviations, respectively, of natural year-to-year variations). The signal of observed temperature change emerged earlier in tropical South America than mid-
North America even though the changes were of a smaller magnitude. (Note that those regions were chosen because of the longer length of their observational
record; see Figure 1.14 for more regions).
FAQ
153
Frequently Asked Questions
FAQ 1.3 | What Can Past Climate Teach Us About the Future?
In the past, the Earth has experienced prolonged periods of elevated greenhouse gas concentrations that caused
global temperatures and sea levels to rise. Studying these past warm periods informs us about the potential long-
term consequences of increasing greenhouse gases in the atmosphere.
Rising greenhouse gas concentrations are driving profound changes to the Earth system, including global
warming, sea level rise, increases in climate and weather extremes, ocean acidification, and ecological shifts
(FAQ 2.2 and FAQ 7.1). The vast majority of instrumental observations of climate began during the 20th century,
when greenhouse gas emissions from human activities became the dominant driver of changes in Earth’s
climate (FAQ 3.1).
As scientists seek to refine our understanding of Earth’s climate system and how it may evolve in coming decades
to centuries, past climate states provide a wealth of insights. Data about these past states help to establish the
relationship between natural climate drivers and the history of changes in global temperature, global sea levels,
the carbon cycle, ocean circulation, and regional climate patterns, including climate extremes. Guided by such
data, scientists use Earth system models to identify the chain of events underlying the transitions between past
climatic states (FAQ 3.3). This is important because during present-day climate change, just as in past climate
changes, some aspects of the Earth system (e.g., surface temperature) respond to changes in greenhouse gases
on a time scale of decades to centuries, while others (e.g., sea level and the carbon cycle) respond over centuries
to millennia (FAQ 5.3). In this way, past climate states serve as critical benchmarks for climate model simulations,
improving our understanding of the sequences, rates, and magnitude of future climate change over the next
decades to millennia.
Analyzing previous warm periods caused by natural factors can help us understand how key aspects of the
climate system evolve in response to warming. For example, one previous warm-climate state occurred roughly
125,000 years ago, during the Last Interglacial period, when slight variations in the Earth’s orbit triggered
a sequence of changes that caused about 1°C–2°C of global warming and about 2–8 m of sea level rise relative
to the 1850–1900, even though atmospheric carbon dioxide concentrations were similar to 1850–1900 values
(FAQ 1.3, Figure 1). Modelling studies highlight that increased summer heating in the higher latitudes of the
FAQ Northern Hemisphere during this time caused widespread melting of snow and ice, reducing the reflectivity of
the planet and increasing the absorption of solar energy by the Earth’s surface. This gave rise to global-scale
warming, which led in turn to further ice loss and sea level rise. These self-reinforcing positive feedback cycles are
a pervasive feature of Earth’s climate system, with clear implications for future climate change under continued
greenhouse gas emissions. In the case of sea level rise, these cycles evolved over several centuries to millennia,
reminding us that the rates and magnitude of sea level rise in the 21st century are just a fraction of the sea level
rise that will ultimately occur after the Earth system fully adjusts to current levels of global warming.
Roughly 3 million years ago, during the Pliocene Epoch, the Earth witnessed a prolonged period of elevated
temperatures (2.5°C–4°C higher than 1850–1900) and higher sea levels (5–25 m higher than 1850–1900), in
combination with atmospheric carbon dioxide concentrations similar to those of the present day. The fact that
Pliocene atmospheric carbon dioxide concentrations were similar to the present, while global temperatures and
sea levels were significantly higher, reflects the difference between an Earth system that has fully adjusted to
changes in natural drivers (the Pliocene) and one where greenhouse gases concentrations, temperature, and sea
level rise are still increasing (present day). Much about the transition into the Pliocene climate state – in terms of
key causes, the role of cycles that hastened or slowed the transition, and the rate of change in climate indicators
such as sea level – remain topics of intense study by climate researchers, using a combination of paleoclimate
observations and Earth system models. Insights from such studies may help to reduce the large uncertainties
around estimates of global sea level rise by 2300, which range from 0.3 m to 3 m above 1850–1900 (in a low-
emissions scenario) to as much as 16 m higher than 1850–1900 (in a very high-emissions scenario that includes
accelerating structural disintegration of the polar ice sheets).
While present-day warming is unusual in the context of the recent geologic past in several different ways
(FAQ 2.1), past warm climate states present a stark reminder that the long-term adjustment to present-day
atmospheric carbon dioxide concentrations has only just begun. That adjustment will continue over the coming
centuries to millennia.
154
Frequently Asked Questions
FAQ 1.3: What can the past tell us about the future?
Past warm periods inform about the potential consequences of rising greenhouse gases in the atmosphere.
1135
ppm
High
+3.3-5.7°C
+2.5-4.0°C
+5-25m
*Triggered by changes in the Earth’s orbit, which redistributed incoming solar energy between seasons and latitudes
FAQ 1.3, Figure 1 | Comparison of past, present and future. Schematic of atmospheric carbon dioxide concentrations, global temperature, and global
sea level during previous warm periods as compared to 1850–1900, present-day (2011–2020), and future (2100) climate change scenarios corresponding to
low-emissions scenarios (SSP1-2.6; lighter colour bars) and very high-emissions scenarios (SSP5-8.5; darker colour bars).
155
Frequently Asked Questions
FAQ 2.1 | The Earth’s Temperature Has Varied Before. How Is the Current Warming Any Different?
Earth’s climate has always changed naturally, but both the global extent and rate of recent warming are unusual.
The recent warming has reversed a slow, long-term cooling trend, and research indicates that global surface
temperature is higher now than it has been for millennia.
While climate can be characterized by many variables, temperature is a key indicator of the overall climate state, and
global surface temperature is fundamental to characterizing and understanding global climate change, including
Earth’s energy budget. A rich variety of geological evidence shows that temperature has changed throughout
Earth’s history. A variety of natural archives from around the planet, such as ocean and lake sediments, glacier
ice and tree rings, shows that there were times in the past when the planet was cooler, and times when it was
warmer. While our confidence in quantifying large-scale temperature changes generally decreases the farther
back in time we look, scientists can still identify at least four major differences between the recent warming and
those of the past.
It’s warming almost everywhere. During decades and centuries of the past 2000 years, some regions warmed
more than the global average while, at the same time, other regions cooled. For example, between the 10th
and 13th centuries, the North Atlantic region warmed more than many other regions. In contrast, the pattern of
recent surface warming is globally more uniform than for other decadal to centennial climate fluctuations over
at least the past two millennia.
It’s warming rapidly. Over the past 2 million years, Earth’s climate has fluctuated between relatively warm
interglacial periods and cooler glacial periods, when ice sheets grew over vast areas of the northern continents.
Intervals of rapid warming coincided with the collapse of major ice sheets, heralding interglacial periods such as
the present Holocene Epoch, which began about 12,000 years ago. During the shift from the last glacial period
to the current interglacial, the total temperature increase was about 5°C. That change took about 5000 years,
with a maximum warming rate of about 1.5°C per thousand years, although the transition was not smooth. In
contrast, Earth’s surface has warmed approximately 1.1°C since 1850–1900. However, even the best reconstruction
of global surface temperature during the last deglacial period is too coarsely resolved for direct comparison with
FAQ a period as short as the past 150 years. But for the past 2000 years, we have higher-resolution records that show
that the rate of global warming during the last 50 years has exceeded the rate of any other 50-year period.
Recent warming reversed a long-term global cooling trend. Following the last major glacial period, global
surface temperature peaked by around 6500 years ago, then slowly cooled. The long-term cooling trend was
punctuated by warmer decades and centuries. These fluctuations were minor compared with the persistent and
prominent warming that began in the mid-19th century when the millennial-scale cooling trend was reversed.
It’s been a long time since it’s been this warm. Averaged over the globe, surface temperatures of the past decade
were probably warmer than when the long cooling trend began around 6500 years ago. If so, we need to look
back to at least the previous interglacial period, around 125,000 years ago, to find evidence for multi-centennial
global surface temperatures that were warmer than now.
Previous temperature fluctuations were caused by large-scale natural processes, while the current warming is
largely due to human causes (see, for example, FAQ 1.3, FAQ 3.1). But understanding how and why temperatures
have changed in the past is critical for understanding the current warming and how human and natural
influences will interact to determine what happens in the future. Studying past climate changes also makes it
clear that, unlike previous climate changes, the effects of recent warming are occurring on top of stresses that
make humans and nature vulnerable to changes in ways that they have never before experienced (for example,
see FAQ 11.2, FAQ12.3).
156
Frequently Asked Questions
It is warming rapidly
FAQ
The warming reversed
a long-term cooling
157
Frequently Asked Questions
The evidence for climate change rests on more than just increasing surface temperatures. A broad range of
indicators collectively leads to the inescapable conclusion that we are witnessing rapid changes to many aspects
of our global climate. We are seeing changes in the atmosphere, ocean, cryosphere, and biosphere. Our scientific
understanding depicts a coherent picture of a warming world.
We have long observed our changing climate. From the earliest scientists taking meteorological observations
in the 16th and 17th centuries to the present, we have seen a revolution in our ability to observe and diagnose
our changing climate. Today we can observe diverse aspects of our climate system from space, from aircraft
and weather balloons, using a range of ground-based observing technologies, and using instruments that can
measure to great depths in the ocean.
Observed changes in key indicators point to warming over land areas. Global surface temperature over land has
increased since the late 19th century, and changes are apparent in a variety of societally relevant temperature
extremes. Since the mid-1950s the troposphere (i.e., the lowest few km of the atmosphere) has warmed, and
precipitation over land has increased. Near-surface specific humidity (i.e., water vapour) over land has increased
since at least the 1970s. Aspects of atmospheric circulation have also evolved since the mid-20th century, including
a poleward shift of mid-latitude storm tracks.
Changes in the global ocean point to warming as well. Global average sea surface temperature has increased
since the late 19th century. The heat content of the global ocean has increased since the 19th century, with more
than 90% of the excess energy accumulated in the climate system being stored in the ocean. This ocean warming
has caused ocean waters to expand, which has contributed to the increase in global sea level in the past century.
The relative acidity of the ocean has also increased since the early 20th century, caused by the uptake of carbon
dioxide from the atmosphere, and oxygen loss is evident in the upper ocean since the 1970s.
Significant changes are also evident over the cryosphere – the portion of the Earth where water is seasonally or
continuously frozen as snow or ice. There have been decreases in Arctic sea ice area and thickness and changes
in Antarctic sea ice extent since the mid-1970s. Spring snow cover in the Northern Hemisphere has decreased
FAQ
since the late-1970s, along with an observed warming and thawing of permafrost (perennially frozen ground).
The Greenland and Antarctic ice sheets are shrinking, as are the vast majority of glaciers worldwide, contributing
strongly to the observed sea level rise.
Many aspects of the biosphere are also changing. Over the last century, long-term ecological surveys show that
many land species have generally moved poleward and to higher elevations. There have been increases in green
leaf area and/or mass (i.e., global greenness) since the early 1980s, and the length of the growing season has
increased over much of the extratropical Northern Hemisphere since at least the mid-20th century. There is also
strong evidence that various phenological metrics (such as the timing of fish migrations) for many marine species
have changed in the last half century.
Change is apparent across many components of the climate system. It has been observed using a very broad range
of techniques and analysed independently by numerous groups around the world. The changes are consistent in
pointing to a climate system that has undergone rapid warming since the industrial revolution.
158
Frequently Asked Questions
AIR
Atmospheric circulation
Surface
temperature Permafrost
Sea level
extent
Ocean
acidification
Global Glaciers
greenness FAQ
Species Growing
range shifts season length
Precipitation
FAQ 2.2, Figure 1 | Synthesis of significant changes observed in the climate system over the past several decades. Upwards, downwards
and circling arrows indicate increases, decreases and changes, respectively. Independent analyses of many components of the climate system that would be
expected to change in a warming world exhibit trends consistent with warming. Note that this list is not comprehensive.
159
Frequently Asked Questions
FAQ 3.1 | How Do We Know Humans Are Responsible for Climate Change?
The dominant role of humans in driving recent climate change is clear. This conclusion is based on a synthesis of
information from multiple lines of evidence, including direct observations of recent changes in Earth’s climate;
analyses of tree rings, ice cores, and other long-term records documenting how the climate has changed in the
past; and computer simulations based on the fundamental physics that governs the climate system.
Climate is influenced by a range of factors. There are two main natural drivers of variations in climate on time
scales of decades to centuries. The first is variations in the sun’s activity, which alter the amount of incoming
energy from the sun. The second is large volcanic eruptions, which increase the number of small particles (aerosols)
in the upper atmosphere that reflect sunlight and cool the surface–an effect that can last for several years
(see also FAQ 3.2). The main human drivers of climate change are increases in the atmospheric concentrations
of greenhouse gases and of aerosols from burning fossil fuels, land use and other sources. The greenhouse
gases trap infrared radiation near the surface, warming the climate. Aerosols, like those produced naturally
by volcanoes, on average cool the climate by increasing the reflection of sunlight. Multiple lines of evidence
demonstrate that human drivers are the main cause of recent climate change.
The current rates of increase of the concentration of the major greenhouse gases (carbon dioxide, methane
and nitrous oxide) are unprecedented over at least the last 800,000 years. Several lines of evidence clearly show
that these increases are the results of human activities. The basic physics underlying the warming effect of
greenhouse gases on the climate has been understood for more than a century, and our current understanding
has been used to develop the latest generation climate models (see FAQ 3.3). Like weather forecasting models,
climate models represent the state of the atmosphere on a grid and simulate its evolution over time based on
physical principles. They include a representation of the ocean, sea ice and the main processes important in
driving climate and climate change.
Results consistently show that such climate models can only reproduce the observed warming (black line in
FAQ 3.1, Figure 1) when including the effects of human activities (grey band in FAQ 3.1, Figure 1), in particular
the increasing concentrations of greenhouse gases. These climate models show a dominant warming effect of
FAQ greenhouse gas increases (red band, which shows the warming effects of greenhouse gases by themselves),
which has been partly offset by the cooling effect of increases in atmospheric aerosols (blue band). By contrast,
simulations that include only natural processes, including internal variability related to El Niño and other similar
variations, as well as variations in the activity of the sun and emissions from large volcanoes (green band in
FAQ 3.1, Figure 1), are not able to reproduce the observed warming. The fact that simulations including only
natural processes show much smaller temperature increases indicates that natural processes alone cannot explain
the strong rate of warming observed. The observed rate can only be reproduced when human influence is added
to the simulations.
Moreover, the dominant effect of human activities is apparent not only in the warming of global surface
temperature, but also in the pattern of warming in the lower atmosphere and cooling in the stratosphere,
warming of the ocean, melting of sea ice, and many other observed changes. An additional line of evidence for
the role of humans in driving climate change comes from comparing the rate of warming observed over recent
decades with that which occurred prior to human influence on climate. Evidence from tree rings and other
paleoclimate records shows that the rate of increase of global surface temperature observed over the past fifty
years exceeded that which occurred in any previous 50-year period over the past 2000 years (see FAQ 2.1).
Taken together, this evidence shows that humans are the dominant cause of observed global warming over
recent decades.
160
Frequently Asked Questions
2.5
Global surface temperature change since 1850
0.5
Natural causes
0
-1.0
-1.5
1850 1900 1950 2000 2020
FAQ 3.1, Figure 1 | Observed warming (1850–2019) is only reproduced in simulations including human influence. Global surface temperature
changes in observations, compared to climate model simulations of the response to all human and natural forcings (grey band), greenhouse gases only
(red band), aerosols and other human drivers only (blue band) and natural forcings only (green band). Solid coloured lines show the multi-model mean,
and coloured bands show the 5–95% range of individual simulations.
FAQ
161
Frequently Asked Questions
FAQ 3.2 | What is Natural Variability and How Has it Influenced Recent Climate Changes?
Natural variability refers to variations in climate that are caused by processes other than human influence.
It includes variability that is internally generated within the climate system and variability that is driven by
natural external factors. Natural variability is a major cause of year-to-year changes in global surface climate and
can play a prominent role in trends over multiple years or even decades. But the influence of natural variability
is typically small when considering trends over periods of multiple decades or longer. When estimated over the
entire historical period (1850 –2020), the contribution of natural variability to global surface warming of –0.23°C
to + 0.23°C is small compared to the warming of about 1.1°C observed during the same period, which has been
almost entirely attributed to the human influence.
Paleoclimatic records (indirect measurements of climate that can extend back many thousands of years) and
climate models all show that global surface temperatures have changed significantly over a wide range of time
scales in the past. One of these reasons is natural variability, which refers to variations in climate that are either
internally generated within the climate system or externally driven by natural changes. Internal natural variability
corresponds to a redistribution of energy within the climate system (for example via atmospheric circulation
changes similar to those that drive the daily weather) and is most clearly observed as regional, rather than
global, fluctuations in surface temperature. External natural variability can result from changes in the Earth’s
orbit, small variations in energy received from the sun, or from major volcanic eruptions. Although large orbital
changes are related to global climate changes of the past, they operate on very long time scales (i.e., thousands
of years). As such, they have displayed very little change over the past century and have had very little influence
on temperature changes observed over that period. On the other hand, volcanic eruptions can strongly cool the
Earth, but this effect is short-lived and their influence on surface temperatures typically fades within a decade
of the eruption.
To understand how much of observed recent climate change has been caused by natural variability (a process
referred to as attribution), scientists use climate model simulations. When only natural factors are used to force
climate models, the resulting simulations show variations in climate on a wide range of time scales in response to
volcanic eruptions, variations in solar activity, and internal natural variability. However, the influence of natural
FAQ climate variability typically decreases as the time period gets longer, such that it only has mild effects on multi-
decadal and longer trends (FAQ 3.2, Figure 1).
Consequently, over periods of a couple of decades or less, natural climate variability can dominate the human-
induced surface warming trend – leading to periods with stronger or weaker warming, and sometimes even
cooling (FAQ 3.2, Figure 1, left and centre). Over longer periods, however, the effect of natural variability is
relatively small (FAQ 3.2, Figure 1, right). For instance, over the entire historical period (1850–2019), natural
variability is estimated to have caused between –0.23°C and +0.23°C of the observed surface warming of
about 1.1°C. This means that the bulk of the warming has been almost entirely attributed to human activities,
particularly emissions of greenhouse gases (FAQ 3.1).
Another way to picture natural variability and human influence is to think of a person walking a dog. The path
of the walker represents the human-induced warming, while their dog represents natural variability. Looking at
global surface temperature changes over short periods is akin to focusing on the dog. The dog sometimes moves
ahead of the owner and other times behind. This is similar to natural variability that can weaken or amplify
warming on the short term. In both cases it is difficult to predict where the dog will be or how the climate will
evolve in the near future. However, if we pull back and focus on the slow steady steps of the owner, the path of
the dog is much clearer and more predictable, as it follows the path of its owner. Similarly, human influence on
the climate is much clearer over longer time periods.
162
Frequently Asked Questions
FAQ 3.2 What is natural variability and how has it influenced recent climate changes?
Natural variability can alter global temperature over short time scales (1 year to ~2 decades) but it has a minimal
influence on longer time scales. Since 1850, natural variability ( ) has caused between -0.23°C and 0.23°C
of global temperature change, compared to the warming of about 1.1°C observed ( ) over that period.
Annual (1 year) variations Decadal (10 year) variations Multi-decadal (30 year) variations
Dominated by natural variability Less influenced by natural variability, Dominated by the human influence
but natural cooling or more intense
warming can still occur
0.8 0.8 0.8
Global surface temperature change (°C)
0.6
0.4
0.2
-0.2
-0.4
-0.6
FAQ 3.2, Figure 1 | Annual (left), decadal (middle) and multi-decadal (right) variations in average global surface temperature. The thick
black line is an estimate of the human contribution to temperature changes, based on climate models, whereas the green lines show the combined effect
of natural variations and human-induced warming, different shadings of green represent different simulations, which can be viewed as showing a range of
potential pasts. The influence of natural variability is shown by the green bars, and it decreases on longer time scales. The data is sourced from the CESM1
large ensemble.
FAQ
163
Frequently Asked Questions
Yes, climate models have improved and continue to do so, becoming better at capturing complex and small-
scale processes and at simulating present-day mean climate conditions. This improvement can be measured by
comparing climate simulations against historical observations. Both the current and previous generations of
models show that increases in greenhouse gases cause global warming. While past warming is well simulated by
the new generation models as a group, some individual models simulate past warming that is either below or
above what is observed. The information about how well models simulate past warming, as well as other insights
from observations and theory, are used to refine this Report’s projections of global warming.
Climate models are important tools for understanding past, present and future climate change. They are
sophisticated computer programs that are based on fundamental laws of physics of the atmosphere, ocean,
ice, and land. Climate models perform their calculations on a three-dimensional grid made of small bricks
or ‘gridcells’ of about 100 km across. Processes that occur on scales smaller than the model grid cells (such
as the transformation of cloud moisture into rain) are treated in a simplified way. This simplification is done
differently in different models. Some models include more processes and complexity than others; some represent
processes in finer detail (smaller grid cells) than others. Hence the simulated climate and climate change vary
between models.
Climate modelling started in the 1950s and, over the years, models have become increasingly sophisticated as
computing power, observations and our understanding of the climate system have advanced. The models used
in the IPCC First Assessment Report published in 1990 correctly reproduced many aspects of climate (FAQ 1.1).
The actual evolution of the climate since then has confirmed these early projections, when accounting for the
differences between the simulated scenarios and actual emissions. Models continue to improve and get better
and better at simulating the large variety of important processes that affect climate. For example, many models
now simulate complex interactions between different aspects of the Earth system, such as the uptake of carbon
dioxide by vegetation on land and by the ocean, and the interaction between clouds and air pollutants. While
some models are becoming more comprehensive, others are striving to represent processes at higher resolution,
for example to better represent the vortices and swirls in currents responsible for much of the transport of heat
FAQ in the ocean.
Scientists evaluate the performance of climate models by comparing historical climate model simulations to
observations. This evaluation includes comparison of large-scale averages as well as more detailed regional and
seasonal variations. There are two important aspects to consider: (i) how models perform individually and (ii)
how they perform as a group. The average of many models often compares better against observations than any
individual model, since errors in representing detailed processes tend to cancel each other out in multi-model
averages.
As an example, FAQ 3.3 Figure 1 compares simulations from the three most recent generations of models (available
around 2008, 2013 and 2021) with observations of three climate variables. It shows the correlation between
simulated and observed patterns, where a value of 1 represents perfect agreement. Many individual models of
the new generation perform significantly better, as indicated by values closer to 1. As a group, each generation
out-performs the previous generation: the multi-model average (shown by the longer lines) is progressively closer
to 1. The vertical extent of the colored bars indicates the range of model performance across each group. The
top of the bar moves up with each generation, indicating improved performance of the best performing models
from one generation to the next. In the case of precipitation, the performance of the worst performing models
is similar in the two most recent model generations, increasing the spread across models.
Developments in the latest generation of climate models, including new and better representation of physical,
chemical and biological processes, as well as higher resolution, have improved the simulation of many aspects
of the Earth system. These simulations, along with the evaluation of the ability of the models to simulate past
warming as well as the updated assessment of the temperature response to a doubling of CO2 in the atmosphere,
are used to estimate the range of future global warming (FAQ 7.3).
164
Frequently Asked Questions
Better model
performance
Individual models
Multi-model
0.9 average
Smaller
spread
Correlation
across
models
0.8
Larger
spread
0.7 across
models
CMIP6
CMIP5
Poorer model CMIP3
performance
0.6
Near-Surface Precipitation Sea Level
Air Temperature Pressure
FAQ
FAQ 3.3, Figure 1 | Pattern correlations between models and observations of three different variables: surface air temperature,
precipitation and sea level pressure. Results are shown for the three most recent generations of models, from the Coupled Model Intercomparison
Project (CMIP): CMIP3 (orange), CMIP5 (blue) and CMIP6 (purple). Individual model results are shown as short lines, along with the corresponding ensemble
average (long line). For the correlations the yearly averages of the models are compared with the reference observations for the period 1980–1999, with
1 representing perfect similarity between the models and observations. CMIP3 simulations performed in 2004-2008 were assessed in the IPCC Fourth
Assessment, CMIP5 simulations performed in 2011–2013 were assessed in the IPCC Fifth Assessment, and CMIP6 simulations performed in 2018–2021 are
assessed in this Report.
165
Frequently Asked Questions
FAQ 4.1 | How Will the Climate Change Over the Next Twenty Years?
The parts of the climate system that have shown clear increasing or decreasing trends in recent decades will
continue these trends for at least the next twenty years. Examples include changes in global surface temperature,
Arctic sea ice cover, and global average sea level. However, over a period as short as twenty years, these trends are
substantially influenced by natural climate variability, which can either amplify or attenuate the trend expected
from the further increase in greenhouse gas concentrations.
Twenty years are a long time by human standards but a short time from a climate point of view. Emissions of
greenhouse gases will continue over the next twenty years, as assumed in all the scenarios considered in this
Report, albeit with varying rates. These emissions will further increase concentrations of greenhouse gases in
the atmosphere (see FAQ 4.2), leading to continued trends in global surface warming and other parts of the
climate system, including Arctic sea ice and global average sea level (see FAQ 9.2). FAQ 4.1, Figure 1 shows that
both global surface temperature rise and the shrinking of sea ice in the Arctic will continue, with little difference
between high- and low-emissions scenarios over the next 20 years (that is, between the red and blue lines).
However, these expected trends will be overlain by natural climate variability (see FAQ 3.2). First, a major volcanic
eruption might occur, such as the 1991 eruption of Mt. Pinatubo on the Philippines; such an eruption might cause
a global surface cooling of a few tenths of a degree Celsius lasting several years. Second, both atmosphere and
ocean show variations that occur spontaneously, without any external influence. These variations range from
localized weather systems to continent- and ocean-wide patterns and oscillations that change over months, years,
or decades. Over a period of twenty years, natural climate variability strongly influences many climate quantities,
when compared to the response to the increase in greenhouse gas concentrations from human activities. The effect
of natural variability is illustrated by the very different trajectories that individual black, red or blue lines can
take in FAQ 4.1, Figure 1. Whether natural variability would amplify or attenuate the human influence cannot
generally be predicted out to twenty years into the future. Natural climate variability over the next twenty years
thus constitutes an uncertainty that at best can be quantified accurately but that cannot be reduced.
Locally, the effect of natural variability would be much larger still. Simulations (not shown here) indicate that,
FAQ locally, a cooling trend over the next twenty years cannot be ruled out, even under the high-emissions scenario –
at a small number of locations on Earth, but these might lie anywhere. Globally, though, temperatures would
rise under all scenarios.
In summary, while the direction of future change is clear for the two important climate quantities shown here –
the global surface temperature and the Arctic sea ice area in September – the magnitude of the change is much
less clear because of natural variability.
166
Frequently Asked Questions
1.5
SSP3–7.0
SSP1–2.6
1.0
Historical
0.5
0.0
-0.5
2000 2010 2020 2030 2040
Low High
Emission
Sea ice area change (millions of km2) scenario
(Arctic – September)
0
FAQ
SSP1–2.6
SSP3–7.0
-2
Historical
-4
-6
2000 2010 2020 2030 2040
Year
FAQ 4.1, Figure 1 | Simulations over the period 1995–2040, encompassing the recent past and the next twenty years, of two important
indicators of global climate change. (Top) Global surface temperature, and (bottom), the area of Arctic sea ice in September. Both quantities are
shown as deviations from the average over the period 1995–2014. The grey curves are for the historical period ending in 2014; the blue curves represent
a low-emissions scenario (SSP1‑2.6) and the red curves one high-emissions scenario (SSP3‑7.0).
167
Frequently Asked Questions
FAQ 4.2 | How Quickly Would We See the Effects of Reducing Carbon Dioxide Emissions?
The effects of substantial reductions in carbon dioxide emissions would not be apparent immediately, and
the time required to detect the effects would depend on the scale and pace of emissions reductions. Under the
lower-emissions scenarios considered in this Report, the increase in atmospheric carbon dioxide concentrations
would slow visibly after about five to ten years, while the slowing down of global surface warming would be
detectable after about twenty to thirty years. The effects on regional precipitation trends would only become
apparent after several decades.
Reducing emissions of carbon dioxide (CO2) – the most important greenhouse gas emitted by human activities –
would slow down the rate of increase in atmospheric CO2 concentration. However, concentrations would only
begin to decrease when net emissions approach zero, that is, when most or all of the CO2 emitted into the
atmosphere each year is removed by natural and human processes (see FAQ 5.1 and FAQ 5.3). This delay between
a peak in emissions and a decrease in concentration is a manifestation of the very long lifetime of CO2 in the
atmosphere; part of the CO2 emitted by humans remains in the atmosphere for centuries to millennia.
Reducing the rate of increase in CO2 concentration would slow down global surface warming within a decade.
But this reduction in the rate of warming would initially be masked by natural climate variability and might
not be detected for a few decades (see FAQ 1.2, FAQ 3.2 and FAQ 4.1). Detecting whether surface warming has
indeed slowed down would thus be difficult in the years right after emissions reductions begin.
The time needed to detect the effect of emissions reductions is illustrated by comparing low- and high-emissions
scenarios (FAQ 4.2, Figure 1). In the low-emissions scenario (SSP1‑2.6), CO2 emissions level off after 2015 and begin
to fall in 2020, while they keep increasing throughout the 21st century in the high-emissions scenario (SSP3‑7.0).
The uncertainty arising from natural internal variability in the climate system is represented by simulating each
scenario ten times with the same climate model but starting from slightly different initial states back in 1850
(thin lines). For each scenario, the differences between individual simulations are caused entirely by simulated
natural internal variability. The average of all simulations represents the climate response expected for a given
scenario. The climate history that would actually unfold under each scenario would consist of this expected
FAQ response combined with the contribution from natural internal variability and the contribution from potential
future volcanic eruptions (the latter effect is not represented here).
FAQ 4.2, Figure 1 shows that the atmospheric CO2 concentrations differ noticeably between the two scenarios
about five to ten years after the emissions have begun to diverge in year 2015. In contrast, the difference in
global surface temperatures between the two scenarios does not become apparent until later – about two to
three decades after the emissions histories have begun to diverge in this example. This time would be longer if
emissions were reduced more slowly than in the low-emissions scenario illustrated here and shorter in the case
of stronger reductions. Detection would take longer for regional quantities and for precipitation changes, which
vary more strongly from natural causes. For instance, even in the low-emissions scenario, the effect of reduced
CO2 emissions would not become visible in regional precipitation until late in the 21st century.
In summary, it is only after a few decades of reducing CO2 emissions that we would clearly see global temperatures
starting to stabilize. By contrast, short-term reductions in CO2 emissions, such as during the COVID-19 pandemic,
do not have detectable effects on either CO2 concentration or global temperature. Only sustained emissions
reductions over decades would have a widespread effect across the climate system.
168
Frequently Asked Questions
80 2015
Divergence
of scenarios SSP3–7.0
60 High
40 Emission
scenario
20
Low
0
SSP1–2.6
700 SSP3–7.0
410
FAQ
500 SSP1–2.6
400
16.5
SSP3–7.0
16.0
15.5
SSP1–2.6
15.0
14.5
FAQ 4.2, Figure 1 | Observing the benefits of emissions reductions. (Top) Carbon dioxide (CO2) emissions, (middle) CO2 concentration in the atmosphere
and (bottom) effect on global surface temperature for two scenarios: a low-emissions scenario (SSP1‑2.6, blue) and a high-emissions scenario (SSP3‑7.0). In the
low-emissions scenario, CO2 emissions begin to decrease in 2020 whereas they keep increasing throughout the 21st century in the high-emissions scenario.
The thick lines are the average of the 10 individual simulations (thin line) for each scenario. Differences between individual simulations reflect natural variability.
169
Frequently Asked Questions
FAQ 4.3 | At a Given Level of Global Warming, What Are the Spatial Patterns of Climate Change?
As the planet warms, climate change does not unfold uniformly across the globe, but some patterns of regional
change show clear, direct and consistent relationships to increases in global surface temperature. The Arctic
warms more than other regions, land areas warm more than the ocean surface, and the Northern Hemisphere
more than the Southern Hemisphere. Precipitation increases over high latitudes, tropics and large parts of
the monsoon regions, but decreases over the subtropics. For cases like these, we can infer the direction and
magnitude of some regional changes – particularly temperature and precipitation changes – for any given level
of global warming.
The intensity of climate change will depend on the level of global warming. It is possible to identify certain
patterns of regional climate change that occur consistently, but increase in amplitude, across increasing levels of
global warming. Such robust spatial patterns of climate change are largely independent of the specific scenario
(and pathway in time) that results in a given level of global warming. That is, as long as different scenarios result
in the same global warming level, irrespective of the time when this level is attained in each scenario, we can
infer the patterns of regional change that would result from this warming. When patterns of changes are robust,
regional consequences can be assessed for all levels of global warming, for all future time periods, and for all
scenarios. Temperature and precipitation show such robust patterns of changes that are particularly striking.
The high latitudes of the Northern Hemisphere are projected to warm the most, by two to four times the level
of global warming – a phenomenon referred to as Arctic amplification (FAQ 4.3, Figure 1, left). Several processes
contribute to this high rate of warming, including increases in the absorption of solar radiation due to the loss
of reflective sea ice and snow in a warmer world. In the Southern Hemisphere, Antarctica is projected to warm
faster than the mid-latitude Southern Ocean, but the Southern Hemisphere high latitudes are projected to
warm at a reduced amplitude compared to the level of global warming (FAQ 4.3, Figure 1, left). An important
reason for the relatively slower warming of the Southern Hemisphere high latitudes is the upwelling of Antarctic
deep waters that drives a large surface heat uptake in the Southern Ocean.
The warming is generally stronger over land than over the ocean, and in the Northern Hemisphere compared to
FAQ the Southern Hemisphere, and with less warming over the central subpolar North Atlantic and the southernmost
Pacific. The differences are the result of several factors, including differences in how land and ocean areas
absorb and retain heat, the fact that there is more land area in the Northern Hemisphere than in the Southern
Hemisphere, and the influence of ocean circulation. In the Southern Hemisphere, robust patterns of relatively
high warming are projected for subtropical South America, southern Africa, and Australia. The relatively strong
warming in subtropical southern Africa arises from strong interactions between soil moisture and temperature
and from increased solar radiation as a consequence of enhanced subsidence.
Precipitation changes are also proportional to the level of global warming (FAQ 4.3, Figure 1, right), although
uncertainties are larger than for the temperature change. In the high latitudes of both the Southern and Northern
Hemispheres, increases in precipitation are expected as the planet continues to warm, with larger changes
expected at higher levels of global warming (FAQ 4.3, Figure 1, right). The same holds true for the projected
precipitation increases over the tropics and large parts of the monsoon regions. General drying is expected
over the subtropical regions, particularly over the Mediterranean, southern Africa and parts of Australia, South
America, and south-west North America, as well as over the subtropical Atlantic and parts of the subtropical
Indian and Pacific Oceans. Increases in precipitation over the tropics and decreases over the subtropics amplify
with higher levels of global warming.
Some regions that are already dry and warm, such as southern Africa and the Mediterranean, are expected to
become progressively drier and drastically warmer at higher levels of global warming.
In summary, climate change will not affect all the parts of the globe evenly. Rather, distinct regional patterns of
temperature and precipitation change can be identified, and these changes are projected to amplify as the level
of global warming increases.
170
Frequently Asked Questions
Warming will be stronger in the Arctic, Precipitation will increase in high latitudes, the tropics
on land and in the Northern Hemisphere and monsoon regions and decrease in the subtropics
+1.5°C
+3.0°C
FAQ 4.3, Figure 1 | Regional changes in temperature (left) and precipitation (right) are proportional to the level of global warming,
FAQ
irrespective of the scenario through which the level of global warming is reached. Surface warming and precipitation change are shown relative
to the 1850–1900 climate, and for time periods over which the globally averaged surface warming is 1.5°C (top) and 3°C (bottom), respectively. Changes
presented here are based on 31 CMIP6 models using the high-emissions scenario SSP3‑7.0.
171
Frequently Asked Questions
FAQ 5.1 | Is the Natural Removal of Carbon From the Atmosphere Weakening?
For decades, about half of the carbon dioxide (CO2) that human activities have emitted to the atmosphere
has been taken up by natural carbon sinks in vegetation, soils and oceans. These natural sinks of CO2 have
thus roughly halved the rate at which atmospheric CO2 concentrations have increased, and therefore slowed
down global warming. However, observations show that the processes underlying this uptake are beginning to
respond to increasing CO2 in the atmosphere and climate change in a way that will weaken nature’s capacity
to take up CO2 in the future. Understanding of the magnitude of this change is essential for projecting how the
climate system will respond to future emissions and emissions reduction efforts.
30
of yearly CO2 emissions staying in the atmosphere
Atmospheric has remained roughly stable at 44% over the last six
20
growth rate decades, despite continuously increasing CO2 emissions
10
from human activities.
Land
CO2 per year
40
FAQ 5.1, Figure 1 | Atmospheric carbon dioxide (CO2) and natural
20 carbon sinks. (Top) Global emissions of CO2 from human activities and
the growth rate of CO2 in the atmosphere; (middle) the net land and
0 ocean CO2 removal (natural sinks); and (bottom) the fraction of CO2
1960 1980 2000 2020 emitted by human activities remaining in atmosphere from 1960 to 2019.
Years Lines are the five years running mean, error bars denote the uncertainty
of the mean estimate. See Table 5.SM.6 for more information on the data
underlying this figure.
172
Frequently Asked Questions
several factors control how much CO2 is captured: the difference in CO2 partial pressure between the atmosphere
and the surface ocean; wind speeds at the ocean surface; the chemical composition of seawater (that is, its
buffering capacity), which affects how much CO2 can be taken up; and the use of CO2 in photosynthesis by
seawater microalgae. The CO2-enriched surface ocean water is transported to the deep ocean in specific zones
around the globe (such as the Northern Atlantic and the Southern Ocean), effectively storing the CO2 away from
the atmosphere for many decades to centuries. The combined effect of warmer surface ocean temperatures on
these processes is to weaken the future ocean CO2 sink.
The ocean carbon sink is better quantified than the land sink, thanks to direct ocean and atmospheric carbon
observations. The land carbon sink is more challenging to monitor globally, because it varies widely, even
regionally. There is currently no direct evidence that the natural sinks are slowing down, because observable
changes in the fraction of human emissions stored on land or in oceans are small compared to year-to-year
and decadal variations of these sinks. Nevertheless, it is becoming more obvious that atmospheric and climate
changes are affecting the processes controlling the land and ocean sinks.
Since the land and ocean sinks respond to the rise in atmospheric CO2 and to human-induced global warming,
the absolute amount of CO2 taken up by land and ocean will be affected by future CO2 emissions. This also implies
that, if countries manage to strongly reduce global CO2 emissions, or even remove CO2 from the atmosphere,
these sinks will take up less CO2 because of the reduced human perturbation of the carbon cycle. Under future
high-warming scenarios, it is expected that the global ocean and land sinks will stop growing in the second
half of the century as climate change increasingly affects them. Thus, the total amount of CO2 emitted to the
atmosphere and the responses of the natural CO2 sinks will both determine what efforts are required to limit
global warming to a certain level (see FAQ 5.4), underscoring how important it is to understand the evolution of
these natural CO2 sinks.
FAQ
173
Frequently Asked Questions
In the Arctic, large amounts of organic carbon are stored in permafrost – ground that remains frozen throughout
the year. If significant areas of permafrost thaw as the climate warms, some of that carbon may be released
into the atmosphere in the form of carbon dioxide or methane, resulting in additional warming. Projections
from models of permafrost ecosystems suggest that future permafrost thaw will lead to some additional
warming – enough to be important, but not enough to lead to a ‘runaway warming’ situation, where permafrost
thaw leads to a dramatic, self-reinforcing acceleration of global warming.
The Arctic is the biggest climate-sensitive carbon pool on Earth, storing twice as much carbon in its frozen soils,
or permafrost, than is currently stored in the atmosphere. As the Arctic region warms faster than anywhere else
on Earth, there are concerns that this warming could release greenhouse gases to the atmosphere and therefore
significantly amplify climate change.
The carbon in the permafrost has built up over thousands of years, as dead plants have been buried and
accumulated within layers of frozen soil, where the cold prevents the organic material from decomposing. As
the Arctic warms and soils thaw, the organic matter in these soils begins to decompose rapidly and return to the
atmosphere as either carbon dioxide or methane, which are both important greenhouse gases. Permafrost can
also thaw abruptly in a given place, due to melting ice in the ground reshaping Arctic landscapes, lakes growing
and draining, and fires burning away insulating surface soil layers. Thawing of permafrost carbon has already
been observed in the Arctic, and climate models project that much of the shallow permafrost (<3 m depth)
throughout the Arctic would thaw under moderate to high amounts of global warming (2°C–4°C).
While permafrost processes are complex, they are beginning to be included in models that represent the
interactions between the climate and the carbon cycle. The projections from these permafrost carbon models show
a wide range in the estimated strength of a carbon–climate vicious circle, from both carbon dioxide and methane,
equivalent to 14–175 billion tonnes of carbon dioxide released per 1°C of global warming. By comparison, in
2019, human activities have released about 40 billion tonnes of carbon dioxide into the atmosphere. This has two
implications. First, the extra warming caused by permafrost thawing is strong enough that it must be considered
FAQ when estimating the total amount of remaining emissions permitted to stabilize the climate at a given level of
global warming (i.e., the remaining carbon budget, see FAQ 5.4). Second, the models do not identify any one
amount of warming at which permafrost thaw becomes a ‘tipping point’ or threshold in the climate system
that would lead to a runaway global warming. However, models do project that emissions would continuously
increase with warming, and that this trend could last for hundreds of years.
Permafrost can also be found in other cold places (e.g., mountain ranges), but those places contain much less
carbon than in the Arctic. For instance, the Tibetan plateau contains about 3% as much carbon as is stored in
the Arctic. There is also concern about carbon frozen in shallow ocean sediments. These deposits are known as
methane hydrates or clathrates, which are methane molecules locked within a cage of ice molecules. They formed
as frozen soils that were flooded when sea levels rose after the last ice age. If these hydrates thaw, they may
release methane that can bubble up to the surface. The total amount of carbon in permafrost-associated
methane hydrates is much less than the carbon in permafrost soils. Global warming takes millennia to penetrate
into the sediments beneath the ocean, which is why these hydrates are still responding to the last deglaciation.
As a result, only a small fraction of the existing hydrates could be destabilised during the coming century. Even
when methane is released from hydrates, most of it is expected to be consumed and oxidised into carbon dioxide
in the ocean before reaching the atmosphere. The most complete modelling of these processes to date suggests
a release to the atmosphere at a rate of less than 2% of current human-induced methane emissions.
Overall, thawing permafrost in the Arctic appears to be an important additional source of heat-trapping gases
to the atmosphere, more so than undersea hydrates. Climate and carbon cycle models are beginning to consider
permafrost processes. While these models disagree on the exact amount of the heat-trapping gases that will be
released into the atmosphere, they agree that: (i) the amount of such gases released from permafrost will increase
with the amount of global warming; and (ii) the warming effect of thawing permafrost is significant enough to
be considered in estimates of the remaining carbon budgets for limiting future warming.
174
Frequently Asked Questions
0 20 40 60 80+
FAQ 5.2, Figure 1 | The Arctic permafrost is a big pool of carbon that is sensitive to climate change. (Left) Quantity of carbon stored in the
permafrost, to 3 m depth (NCSCDv2 dataset) and (right) area of permafrost vulnerable to abrupt thaw (Circumpolar Thermokarst Landscapes dataset).
FAQ
175
Frequently Asked Questions
FAQ 5.3 | Could Climate Change Be Reversed By Removing Carbon Dioxide From the Atmosphere?
0.5
FAQ 5.3, Figure 1 | Changes in aspects of climate change
0.4 in response to a peak and decline in the atmospheric CO2
0.3
concentration (top panel). The vertical grey dashed line indicates the
Ocean thermal
time of peak CO2 concentration in all panels. It shows that the reversal
expansion (m)
0.2 of global surface warming lags the decrease in the atmospheric CO2
concentration by a few years, the reversal of permafrost area decline
0.1 lags the decrease in atmospheric CO2 by decades, and ocean thermal
expansion continues for several centuries. The quantitative information in
0.0
the figure (i.e., numbers on vertical axes) is not to be emphasized as it
1900 2000 2100 2200 2300
Year results from simulations with just one model and will be different for other
models. The qualitative behaviour, however, can be expected to be largely
model independent.
176
Frequently Asked Questions
if CO2 release equals removal, the atmospheric CO2 concentration would stabilize; and if CO2 removal exceeds
release, the CO2 concentration would decline. This applies in the same way to net CO2 emissions – that is, the sum
of human-caused releases and deliberate removals.
If the CO2 concentration in the atmosphere starts to go down, the Earth’s climate would respond to this change
(FAQ 5.3, Figure 1). Some parts of the climate system take time to react to a change in CO2 concentration, so
a decline in atmospheric CO2 as a result of net negative emissions would not lead to immediate reversal of
all climate change trends. Recent studies have shown that global surface temperature starts to decline within
a few years following a decline in atmospheric CO2, although the decline would not be detectable for decades
due to natural climate variability (see FAQ 4.2). Other consequences of human-induced climate change, such as
reduction in permafrost area, would take decades; yet others, such as warming, acidification and oxygen loss
of the deep ocean, would take centuries to reverse following a decline in the atmospheric CO2 concentration.
Sea level would continue to rise for many centuries to millennia, even if large deliberate CO2 removals were
successfully implemented.
‘Overshoot’ scenarios are a class of future scenarios that are receiving increasing attention, particularly in
the context of ambitious climate goals, such as the global warming limits of 1.5°C or 2°C included in the Paris
Agreement. In these scenarios, a slow rate of reduction in emissions in the near term is compensated by net
negative CO2 emissions in the later part of this century, which results in a temporary breach or ‘overshoot’ of
a given warming level. Due to the delayed reaction of several climate system components, it follows that the
temporary overshoot would result in additional climate changes compared to a scenario that reaches the goal
without overshoot. These changes would take decades to many centuries to reverse, with the reversal taking
longer for scenarios with larger overshoot.
Removing more CO2 from the atmosphere than is emitted into it would indeed begin to reverse some aspects
of climate change, but some changes would still continue in their current direction for decades to millennia.
Approaches capable of large-scale removal of CO2 are still in the state of research and development or unproven
at the scales of deployment necessary to achieve a net reduction in atmospheric CO2 levels. CO2 removal
approaches, particularly those deployed on land, can have undesired side effects on water, food production
and biodiversity. FAQ
177
Frequently Asked Questions
There are several types of carbon budgets. Most often, the term refers to the total net amount of carbon dioxide
(CO2) that can still be emitted by human activities while limiting global warming to a specified level (e.g., 1.5°C
or 2°C above pre-industrial levels). This is referred to as the ‘remaining carbon budget’. Several choices and value
judgements have to be made before it can be unambiguously estimated. When the remaining carbon budget is
combined with all past CO2 emissions to date, a ‘total carbon budget’ compatible with a specific global warming
limit can also be defined. A third type of carbon budget is the ‘historical carbon budget’, which is a scientific way
to describe all past and present sources and sinks of CO2.
The term remaining carbon budget is used to describe the total net amount of CO2 that human activities can
still release into the atmosphere while keeping global warming to a specified level, like 1.5°C or 2°C relative
to pre-industrial temperatures. Emissions of CO2 from human activities are the main cause of global warming.
A remaining carbon budget can be defined because of the specific way CO2 behaves in the Earth system. That
is, global warming is roughly linearly proportional to the total net amount of CO2 emissions that are released
into the atmosphere by human activities – also referred to as cumulative anthropogenic CO2 emissions. Other
greenhouse gases behave differently and have to be accounted for separately.
The concept of a remaining carbon budget implies that, to stabilize global warming at any particular level,
global emissions of CO2 need to be reduced to net zero levels at some point. ‘Net zero CO2 emissions’ describes
a situation where all the anthropogenic emissions of CO2 are counterbalanced by deliberate anthropogenic
removals so that, on average, no CO2 is added or removed from the atmosphere by human activities. Atmospheric
CO2 concentrations in such a situation would gradually decline to a long-term stable level as excess CO2 in the
atmosphere is taken up by ocean and land sinks (see FAQ 5.1). The concept of a remaining carbon budget also
means that, if CO2 emissions reductions are delayed, deeper and faster reductions are needed later to stay within
the same budget. If the remaining carbon budget is exceeded, this will result in either higher global warming
or a need to actively remove CO2 from the atmosphere to reduce global temperatures back down to the desired
level (see FAQ 5.3).
FAQ Estimating the size of remaining carbon budgets depends on a set of choices. These choices include: (1) the global
warming level that is chosen as a limit (for example, 1.5°C or 2°C relative to pre-industrial levels); (2) the probability
with which we want to ensure that warming is held below that limit (for example, a one-in-two, two-in-three,
or higher chance), and (3) how successful we are in limiting emissions of other greenhouse gases that affect the
climate, such as methane or nitrous oxide. These choices can be informed by science, but ultimately represent
subjective choices. Once these choices have been made, to estimate the remaining carbon budget for a given
temperature goal, we can combine knowledge about: how much our planet has warmed already; the amount
of warming per cumulative tonne of CO2; and the amount of warming that is still expected once global net
CO2 emissions are brought down to zero. For example, to limit global warming to 1.5°C above pre-industrial
levels with either a one-in-two (50%) or two-in-three (67%) chance, the remaining carbon budgets amount
to 500 and 400 billion tonnes of CO2, respectively, from 1 January 2020 onward (FAQ 5.4, Figure 1). Currently,
human activities are emitting around 40 billion tonnes of CO2 into the atmosphere in a single year.
The remaining carbon budget depends on how much the world has already warmed to date. This past warming
is caused by historical emissions, which are estimated by looking at the historical carbon budget – a scientific way
to describe all past and present sources and sinks of CO2. It describes how the CO2 emissions from human activities
have redistributed across the various CO2 reservoirs of the Earth system. These reservoirs are the ocean, the land
vegetation, and the atmosphere (into which CO2 was emitted). The share of CO2 that is not taken up by the
ocean or the land, and that thus increases the concentration of CO2 in the atmosphere, causes global warming.
The historical carbon budget tells us that, of the about 2560 billion tonnes of CO2 that were released into the
atmosphere by human activities between the years 1750 and 2019, about a quarter were absorbed by the ocean
(causing ocean acidification) and about a third by the land vegetation. About 45% of these emissions remain in
the atmosphere (see FAQ 5.1). Adding these historical CO2 emissions to estimates of remaining carbon budgets
allows an estimate of the total carbon budget consistent with a specific global warming level.
178
Frequently Asked Questions
In summary, determining a remaining carbon budget – that is, how much CO2 can be released into the atmosphere
while stabilizing global temperature below a chosen level – is well understood but relies on a set of choices.
However, it is clear that, for limiting warming below 1.5°C or 2°C, the remaining carbon budget from 2020
onwards is much smaller than the total CO2 emissions released to date.
Historical budget
GtCO2 already emitted between 1750–2019
2560
GtCO2
FAQ 5.4, Figure 1 | Various types of carbon budgets. Historical cumulative carbon dioxide (CO2) emissions determine to a large degree how much FAQ
the world has warmed to date, while the remaining carbon budget indicates how much CO2 could still be emitted while keeping warming below specific
temperature thresholds. Several factors limit the precision with which the remaining carbon budget can be estimated. Therefore, estimates need to specify the
probability with which they aim at limiting warming to the intended target level (e.g., limiting warming to 1.5°C with a 67% probability).
179
Frequently Asked Questions
FAQ 6.1 | What Are Short-lived Climate Forcers and How Do They Affect the Climate?
Short-lived climate forcers (SLCFs) are compounds such as methane and sulphate aerosols that warm or cool the
Earth’s climate over shorter time scales – from days to years – than greenhouse gases like carbon dioxide, whose
climatic effect lasts for decades, centuries or more. Because SLCFs do not remain in the atmosphere for very
long, their effects on the climate are different from one region to another and can change rapidly in response
to changes in SLCF emissions. As some SLCFs also negatively affect air quality, measures to improve air quality
have resulted in sharp reductions in emissions and concentrations of those SLCFs in many regions over the few
last decades.
The SLCFs include gases as well as tiny particles called aerosols, and they can have a warming or cooling effect
on the climate (FAQ 6.1, Figure 1). Warming SLCFs are either greenhouse gases (e.g., ozone or methane) or
particles like black carbon (also known as soot), which warm the climate by absorbing energy and are sometimes
referred to as short-lived climate pollutants. Cooling SLCFs, on the other hand, are mostly made of aerosol
particles (e.g., sulphate, nitrate and organic aerosols) that cool down the climate by reflecting away more
incoming sunlight.
Some SLCFs do not directly affect the climate but produce climate-active compounds and are referred to as
precursors. SLCFs are emitted both naturally and as a result of human activities, such as agriculture or extraction
of fossil fuels. Many of the human sources, particularly those involving combustion, produce SLCFs at the
same time as carbon dioxide and other long-lived greenhouse gases. Emissions have increased since the start
of industrialization, and humans are now the dominant source for several SLCFs and SLCF precursors, such as
sulphur dioxide (which produces sulphate aerosols) and nitrogen oxides (which produce nitrate aerosols and
ozone), despite strong reductions over the last few decades in some regions due to efforts to improve air quality.
The climatic effect of a chemical compound in the atmosphere depends on two things: (i) how effective it
is at cooling or warming the climate (its radiative efficiency) and (ii) how long it remains in the atmosphere
(its lifetime). Because they have high radiative efficiencies, SLCFs can have a strong effect on the climate even
though they have relatively short lifetimes of up to about two decades after emission. Today, there is a balance
FAQ between warming and cooling from SLCFs, but this can change in the future.
The short lifetime of SLCFs constrains their effects in both space and time. First, of all the SLCFs, methane and
the short-lived halocarbons persist the longest in the atmosphere: up to two decades (FAQ 6.1, Figure 1). This is
long enough to mix in the atmosphere and to spread globally. Most other SLCFs only remain in the atmosphere
for a few days to weeks, which is generally too short for mixing in the atmosphere, sometimes even regionally.
As a result, the SLCFs are unevenly distributed and their effects on the climate are more regional than those of
longer-lived gases. Second, rapid (but sustained) changes in emissions of SLCFs can result in rapid climatic effects.
In addition to the direct warming and cooling effects, SLCFs have many other consequences for the climate system
and for air quality (see FAQ 6.2). For instance, deposition of black carbon on snow darkens its surface, which
subsequently absorbs more solar energy, leading to more melting and more warming. Aerosols also modify the
properties of clouds, which has indirect cooling effects on the climate and causes changes in local rainfall (see
FAQ 7.2). Climate models indicate that SLCFs have altered atmospheric circulation on local and even hemispheric
scales (e.g., monsoons) as well as regional precipitation. For instance, recent observations show that regional
weather is influenced by strong regional contrasts in the evolution of aerosol concentrations, particularly over
South and East Asia.
Although policies to limit climate change and discussions of the so-called remaining carbon budgets primarily
focus on carbon dioxide (see FAQ 5.4), SLCFs can significantly affect temperature changes. It is therefore important
to understand how SLCFs work and to quantify their effects. Because reducing some of the SLCF emissions, such as
methane, can simultaneously reduce warming effects and adverse effects on air quality as well as help attaining
Sustainable Development Goals, mitigation of SLCFs is often viewed as a favourable ‘win-win’ policy option.
180
Frequently Asked Questions
FAQ 6.1: What are short-lived climate forcers and how do they affect the climate?
Short lived climate forcers do not remain for very long in the atmosphere, thus an increase or decrease in their emissions
rapidly affects the climate system.
Methane
(CH4) 12 years
Ozone
(O3)
Weeks
Black
Carbon Days
Nitrate
FAQ
Sulphate Days
Organic
aerosols
FAQ 6.1, Figure 1 | Main short-lived climate forcers, their sources, how long they exist in the atmosphere, and their relative contribution
to global surface temperature changes between 1750 and 2019 (area of the globe). By definition this contribution depends on the lifetime, the
warming/cooling potential (radiative efficiency), and the emissions of each compound in the atmosphere. Blue indicates cooling and orange indicates warming.
Note that, between 1750 and 2019, the cooling contribution from aerosols (blue diamonds and globe) was approximately half the warming contribution from
carbon dioxide.
181
Frequently Asked Questions
FAQ 6.2 | What Are the Links Between Limiting Climate Change and Improving
Air Quality?
Climate change and air quality are intimately linked. Many of the human activities that produce long-lived
greenhouse gases also emit air pollutants, and many of these air pollutants are also ‘short-lived climate forcers’
that affect the climate. Therefore, many options for improving air quality may also serve to limit climate change
and vice versa. However, some options for improving air quality cause additional climate warming, and some
actions that address climate change can worsen air quality.
Climate change and air pollution are both critical environmental issues that are already affecting humanity.
In 2016, the World Health Organization attributed 4.2 million deaths worldwide every year to ambient (outdoor)
air pollution. Meanwhile, climate change impacts water resources, food production, human health, extreme
events, coastal erosion, wildfires, and many other phenomena.
Most human activities, including energy production, agriculture, transportation, industrial processes, waste
management and residential heating and cooling, result in emissions of gaseous and particulate pollutants that
modify the composition of the atmosphere, leading to degradation of air quality as well as to climate change.
These air pollutants are also short-lived climate forcers – substances that affect the climate but remain in the
atmosphere for shorter periods (days to decades) than long-lived greenhouse gases like carbon dioxide (see
FAQ 6.1). While this means that the issues of air pollution and climate change are intimately connected, air
pollutants and greenhouse gases are often defined, investigated and regulated independently of one another in
both the scientific and policy arenas.
Many sources simultaneously emit carbon dioxide and air pollutants. When we drive our fossil fuel vehicles
or light a fire in the fireplace, it is not just carbon dioxide or air pollutants that are emitted, but always both.
It is therefore not possible to separate emissions into two clearly distinct groups. As a result, policies aiming at
addressing climate change may have benefits or side effects for air quality, and vice versa.
For example, some short-term ‘win-win’ policies that simultaneously improve air quality and limit climate change
include the implementation of energy efficiency measures, methane capture and recovery from solid-waste
FAQ management and the oil and gas industry, zero-emissions vehicles, efficient and clean stoves for heating and
cooking, filtering of soot (particulate matter) for diesel vehicles, cleaner brick-kiln technology, practices that
reduce burning of agricultural waste, and the eradication of burning of kerosene for lighting.
There are, however, also ‘win-lose’ actions. For example, wood burning is defined as carbon neutral because
a tree accumulates the same amount of carbon dioxide throughout its lifetime as is released when wood from
that tree is burned. However, burning wood can also result in significant emissions of air pollutants, including
carbon monoxide, nitrogen oxides, volatile organic compounds, and particulate matter, that locally or regionally
affect the climate, human health and ecosystems (FAQ 6.2, Figure 1). Alternatively, decreasing the amount of
sulphate aerosols produced by power and industrial plants, and from maritime transport, improves air quality
but results in a warming influence on the climate, because those sulphate aerosols contribute to cooling the
atmosphere by blocking incoming sunlight.
Air quality and climate change represent two sides of the same coin, and addressing both issues together could
lead to significant synergies and economic benefits while avoiding policy actions that mitigate one of the two
issues but worsen the other.
182
Frequently Asked Questions
Reduce Reduce
Nitrous Carbon
Methane monoxide Nitrate
oxide
(CH4) (CO)
(N2O)
se
ea
Organic
In
Surface
Tropospheric
particulate
Ozone (O3) matter
FAQ
Climate benefit Crop benefits Health benefits Climate penalty
FAQ 6.2, Figure 1 | Links between actions aiming to limit climate change and actions to improve air quality. Greenhouse gases (GHGs) and
aerosols (orange and blue) can affect climate directly. Air pollutants (bottom) can affect human health, ecosystems and climate. All these compounds have
common sources and sometimes interact with each other in the atmosphere which makes it impossible to consider them separately (dotted grey arrows).
183
Frequently Asked Questions
FAQ 7.1 | What Is the Earth’s Energy Budget, and What Does It Tell Us About Climate Change?
The Earth’s energy budget describes the flow of energy within the climate system. Since at least 1970 there
has been a persistent imbalance in the energy flows that has led to excess energy being absorbed by the
climate system. By measuring and understanding these energy flows and the role that human activities play in
changing them, we are better able to understand the causes of climate change and project future climate change
more accurately.
Our planet receives vast amounts of energy every day in the form of sunlight. Around a third of the sunlight is
reflected back to space by clouds, by tiny particles called aerosols, and by bright surfaces such as snow and ice.
The rest is absorbed by the ocean, land, ice and atmosphere. The planet then emits energy back out to space in
the form of thermal radiation. In a world that was not warming or cooling, these energy flows would balance.
Human activity has caused an imbalance in these energy flows.
We measure the influence of various human and natural factors on the energy flows at the top of our atmosphere
in terms of radiative forcings, where a positive radiative forcing has a warming effect and a negative radiative
forcing has a cooling effect. In response to these forcings, the Earth system will either warm or cool, so as to
restore balance through changes in the amount of outgoing thermal radiation (the warmer the Earth, the more
radiation it emits). Changes in Earth’s temperature in turn lead to additional changes in the climate system
(known as climate feedbacks) that either amplify or dampen the original effect. For example, Arctic sea ice has
been melting as the Earth warms, reducing the amount of reflected sunlight and adding to the initial warming
(an amplifying feedback). The most uncertain of those climate feedbacks are clouds, as they respond to warming in
complex ways that affect both the emission of thermal radiation and the reflection of sunlight. However, we are
now more confident that cloud changes, taken together, will amplify climate warming (see FAQ 7.2).
Human activities have unbalanced these energy flows in two main ways. First, increases in greenhouse gas levels
have led to more of the emitted thermal radiation being absorbed by the atmosphere, instead of being released to
space. Second, increases in pollutants have increased the amount of aerosols such as sulphates in the atmosphere
(see FAQ 6.1). This has led to more incoming sunlight being reflected away, by the aerosols themselves and
FAQ through the formation of more cloud drops, which increases the reflectivity of clouds (see FAQ 7.2).
Altogether, the global energy flow imbalance since the 1970s has been just over half a watt per square metre
of the Earth’s surface. This sounds small, but because the imbalance is persistent and because Earth’s surface is
large, this adds up to about 25 times the total amount of primary energy consumed by human society, compared
over 1971 to 2018. Compared to the IPCC Fifth Assessment Report (AR5), we are now better able to quantify
and track these energy flows from multiple lines of evidence, including satellite data, direct measurements of
ocean temperatures, and a wide variety of other Earth system observations (see FAQ 1.1). We also have a better
understanding of the processes contributing to this imbalance, including the complex interactions between
aerosols, clouds and radiation.
Research has shown that the excess energy since the 1970s has mainly gone into warming the ocean (91%),
followed by the warming of land (5%) and the melting of ice sheets and glaciers (3%). The atmosphere has
warmed substantially since 1970, but because it is comprised of thin gases it has absorbed only 1% of the excess
energy (FAQ 7.1, Figure 1). As the ocean has absorbed the vast majority of the excess energy, especially within
its top two kilometres, the deep ocean is expected to continue to warm and expand for centuries to millennia,
leading to long-term sea level rise – even if atmospheric greenhouse gas levels were to decline (see FAQ 5.3).
This is in addition to the sea level rise expected from melting ice sheets and glaciers.
Understanding the Earth’s energy budget al.o helps to narrow uncertainty in future projections of climate.
By testing climate models against what we know about the Earth’s energy budget, we can make more confident
projections of surface temperature changes we might expect this century and beyond.
184
Frequently Asked Questions
Atmosphere 1%
91% 5% Land
Ocean
3%
Ice
Excess energy accumulating
FAQ 7.1, Figure 1 | The Earth’s energy budget compares the flows of incoming and outgoing energy that are relevant for the climate
system. Since at least the 1970s, less energy is flowing out than is flowing in, which leads to excess energy being absorbed by the ocean, land, ice and
atmosphere, with the ocean absorbing 91%.
FAQ
185
Frequently Asked Questions
One of the biggest challenges in climate science has been to predict how clouds will change in a warming world
and whether those changes will amplify or partially offset the warming caused by increasing concentrations of
greenhouse gases and other human activities. Scientists have made significant progress over the past decade and
are now more confident that changes in clouds will amplify, rather than offset, global warming in the future.
Clouds cover roughly two-thirds of the Earth’s surface. They consist of small droplets and/or ice crystals, which
form when water vapour condenses or deposits around tiny particles called aerosols (such as salt, dust, or smoke).
Clouds play a critical role in the Earth’s energy budget at the top of our atmosphere and therefore influence
Earth’s surface temperature (see FAQ 7.1). The interactions between clouds and the climate are complex and
varied. Clouds at low altitudes tend to reflect incoming solar energy back to space, creating a cooling effect by
preventing this energy from reaching and warming the Earth. On the other hand, higher clouds tend to trap
(i.e., absorb and then emit at a lower temperature) some of the energy leaving the Earth, leading to a warming
effect. On average, clouds reflect back more incoming energy than the amount of outgoing energy they trap,
resulting in an overall net cooling effect on the present climate. Human activities since the pre-industrial era
have altered this climate effect of clouds in two different ways: by changing the abundance of the aerosol
particles in the atmosphere and by warming the Earth’s surface, primarily as a result of increases in greenhouse
gas emissions.
The concentration of aerosols in the atmosphere has markedly increased since the pre-industrial era, and this
has had two important effects on clouds. First, clouds now reflect more incoming energy because cloud droplets
have become more numerous and smaller. Second, smaller droplets may delay rain formation, thereby making
the clouds last longer, although this effect remains uncertain. Hence, aerosols released by human activities have
had a cooling effect, counteracting a considerable portion of the warming caused by increases in greenhouse
gases over the last century (see FAQ 3.1). Nevertheless, this cooling effect is expected to diminish in the future, as
air pollution policies progress worldwide, reducing the amount of aerosols released into the atmosphere.
Since the pre-industrial period, the Earth’s surface and atmosphere have warmed, altering the properties of
FAQ clouds, such as their altitude, amount and composition (water or ice), thereby affecting the Earth’s energy budget
and, in turn, changing temperature. This cascading effect of clouds, known as the cloud feedback, could either
amplify or offset some of the future warming and has long been the biggest source of uncertainty in climate
projections. The problem stems from the fact that clouds can change in many ways and that their processes occur
on much smaller scales than global climate models can explicitly represent. As a result, global climate models
have disagreed on how clouds, particularly over the subtropical ocean, will change in the future and whether the
change will amplify or suppress the global warming.
Since the last IPCC Report in 2013 (the Fifth Assessment Report, or AR5), understanding of cloud processes has
advanced with better observations, new analysis approaches and explicit high-resolution numerical simulation of
clouds. Also, current global climate models simulate cloud behaviour better than previous models, due both to
advances in computational capabilities and process understanding. Altogether, this has helped to build a more
complete picture of how clouds will change as the climate warms (FAQ 7.2, Figure 1). For example, the amount
of low-clouds will reduce over the subtropical ocean, leading to less reflection of incoming solar energy, and
the altitude of high-clouds will rise, making them more prone to trapping outgoing energy; both processes
have a warming effect. In contrast, clouds in high latitudes will be increasingly made of water droplets rather
than ice crystals. This shift from fewer, larger ice crystals to smaller but more numerous water droplets will
result in more of the incoming solar energy being reflected back to space and produce a cooling effect. Better
understanding of how clouds respond to warming has led to more confidence than before that future changes
in clouds will, overall, cause additional warming (i.e., by weakening the current cooling effect of clouds). This is
called a positive net cloud feedback.
In summary, clouds will amplify rather than suppress the warming of the climate system in the future, as more
greenhouse gases and fewer aerosols are released to the atmosphere by human activities.
186
Frequently Asked Questions
Present climate Future climate Present climate Future climate Present climate Future climate
Incoming
solar
energy
Incoming
solar
energy
Outgoing
energy
Surface
FAQ 7.2, Figure 1 | Interactions between clouds and the climate, today and in a warmer future. Global warming is expected to alter the altitude
(left) and the amount (centre) of clouds, which will amplify warming. On the other hand, cloud composition will change (right), offsetting some of the
warming. Overall, clouds are expected to amplify future warming.
FAQ
187
Frequently Asked Questions
FAQ 7.3 | What Is Equilibrium Climate Sensitivity and How Does It Relate to Future Warming?
For a given future scenario, climate models project a range of changes in global surface temperature. This range
is closely related to equilibrium climate sensitivity, or ECS, which measures how climate models respond to
a doubling of carbon dioxide in the atmosphere. Models with high climate sensitivity project stronger future
warming. Some climate models of the new generation are more sensitive than the range assessed in the IPCC
Sixth Assessment Report. This leads to end-of-century global warming in some simulations of up to 2°C–3°C
above the current IPCC best estimate. Although these higher warming levels are not expected to occur, high-ECS
models are useful for exploring low-likelihood, high-impact futures.
The equilibrium climate sensitivity (ECS) is defined as the long-term global warming caused by a doubling of
carbon dioxide above its pre-industrial concentration. For a given emissions scenario, much of the uncertainty in
projections of future warming can be explained by the uncertainty in ECS (FAQ 7.3, Figure 1). The significance
of equilibrium climate sensitivity has long been recognized, and the first estimate was presented by Swedish
scientist Svante Arrhenius in 1896.
This Sixth Assessment Report concludes that there is a 90% or more chance (very likely) that the ECS is between
2°C and 5°C. This represents a significant reduction in uncertainty compared to the Fifth Assessment Report,
which gave a 66% chance (likely) of ECS being between 1.5°C and 4.5°C. This reduction in uncertainty has been
possible not through a single breakthrough or discovery but instead by combining evidence from many different
sources and by better understanding their strengths and weaknesses.
• The self-reinforcing processes, called feedback loops, that amplify or dampen the warming in response to
increasing carbon dioxide are now better understood. For example, warming in the Arctic melts sea ice,
resulting in more open ocean area, which is darker and therefore absorbs more sunlight, further intensifying
the initial warming. It remains challenging to represent realistically all the processes involved in these
feedback loops, particularly those related to clouds (see FAQ 7.2). Such identified model errors are now taken
into account, and other known, but generally weak, feedback loops that are typically not included in models
FAQ are now included in the assessment of ECS.
• Historical warming since early industrialisation provides strong evidence that climate sensitivity is not small.
Since 1850, the concentrations of carbon dioxide and other greenhouse gases have increased, and as a result
the Earth has warmed by about 1.1°C. However, relying on this industrial-era warming to estimate ECS is
challenging, partly because some of the warming from greenhouse gases was offset by cooling from aerosol
particles and partly because the ocean is still responding to past increases in carbon dioxide.
• Evidence from ancient climates that had reached equilibrium with greenhouse gas concentrations, such as the
coldest period of the last ice age around 20,000 years ago, or warmer periods further back in time, provide
useful data on the ECS of the climate system (see FAQ 1.3).
• Statistical approaches linking model ECS values with observed changes, such as global warming since the
1970s, provide complementary evidence.
All four lines of evidence rely, to some extent, on climate models, and interpreting the evidence often benefits
from model diversity and spread in modelled climate sensitivity. Furthermore, high-sensitivity models can provide
important insights into futures that have a low likelihood of occurring but that could result in large impacts.
But, unlike in previous assessments, climate models are not considered a line of evidence in their own right in the
IPCC Sixth Assessment Report.
The ECS of the latest climate models is, on average, higher than that of the previous generation of models
and also higher than this Report’s best estimate of 3.0°C. Furthermore, the ECS values in some of the new
models are both above and below the 2°C to 5°C very likely range, and although such models cannot be ruled
out as implausible solely based on their ECS, some simulations display climate change that is inconsistent with
the observed changes when tested with ancient climates. A slight mismatch between models and this Report’s
assessment is only natural because this Report’s assessment is largely based on observations and an improved
understanding of the climate system.
188
Frequently Asked Questions
6 6
5 5
IPCC best
2
estimate
(and range)
4 4
1
IPCC best 3
3 estimate 3
(and range)
2 2
FAQ 7.3, Figure 1 | Equilibrium climate sensitivity and future warming. (left) Equilibrium climate sensitivities for the current generation (Coupled
Model Intercomparison Project Phase 6, CMIP6) climate models, and the previous (CMIP5) generation. The assessed range in this Report (AR6) is also shown.
(right) Climate projections of CMIP5, CMIP6 and AR6 for the very high-emissions scenarios RCP8.5, and SSP5‑8.5, respectively. The thick horizontal lines FAQ
represent the multi-model average and the thin horizontal lines represent the results of individual models. The boxes represent the model ranges for CMIP5
and CMIP6 and the range assessed in AR6.
189
Frequently Asked Questions
FAQ 8.1 | How Does Land Use Change Alter the Water Cycle?
The ways in which humans use and change land cover, for example by converting fields to urban areas or clearing
forests, can affect every aspect of the water cycle. Land-use changes can alter precipitation patterns and how
water is absorbed into the ground, flows into streams and rivers, or floods the land surface, as well as how
moisture evaporates back into the air. Changes in any of these aspects of the interconnected water cycle can
affect the entire cycle and the availability of freshwater resources.
Land use describes the combination of activities and ground cover defining each area of the Earth’s continental
surface. Altering land use can modify the exchange of water between the atmosphere, soil and subsurface
(FAQ 8.1, Figure 1).
For instance, changes in land cover can affect the ability of soils to soak up surface water (infiltration). When soil
loses its capacity to soak up water, precipitation that would normally infiltrate and contribute to groundwater
reserves will instead overflow, increasing surface water (runoff) and the likelihood of flooding. For example,
changing from vegetation to urban cover can cause water to flow rapidly over buildings, roads and driveways
and into drains rather than soaking into the ground. Deforestation over wide areas can also directly reduce
soil moisture, evaporation and rainfall locally but can also cause regional temperature changes that affect
rainfall patterns.
Extracting water from the ground and river systems for agriculture, industry and drinking water depletes
groundwater and can increase surface evaporation because water that was previously in the ground is now in
direct contact with the atmosphere, being available for evaporation.
Changing land use can also alter how wet the soil is, influencing how quickly the ground heats up and cools
down and the local water cycle. Drier soils evaporate less water into the air but heat up more in the day. This can
lead to warmer, more buoyant plumes of air that can promote cloud development and precipitation if there is
enough moisture in the air.
Changes in land use can also modify the amount of tiny aerosol particles in the air. For instance, industrial and
FAQ domestic activities can contribute to aerosol emissions, as do natural environments such as forests or salt lakes.
Aerosols cool down global temperature by blocking out sunlight but can also affect the formation of clouds and
therefore the occurrence of precipitation (see FAQ 7.2).
Vegetation plays an important role in soaking up soil moisture and evaporating water into the air (transpiration)
through tiny holes (stomata) that allow the plants to take in carbon dioxide. Some plants are better at retaining
water than others, so changes in vegetation can affect how much water infiltrates into the ground, flows into
streams and rivers, or is evaporated.
More globally, land-use change is currently responsible for about 15% of the emissions of carbon dioxide
from human activities, leading to global warming, which in turn affects precipitation, evaporation, and plant
transpiration. In addition, higher atmospheric concentrations of carbon dioxide due to human activities can
make plants more efficient at retaining water because the stomata do not need to open so widely. Improved
land and water management (e.g., reforestation, sustainable irrigation) can also contribute to reducing climate
change and adapting to some of its adverse consequences.
In summary, there is abundant evidence that changes in land use and land cover alter the water cycle globally,
regionally and locally, by changing precipitation, evaporation, flooding, groundwater, and the availability of
freshwater for a variety of uses. Since all the components of the water cycle are connected (and linked to the
carbon cycle), changes in land use trickle down to many other components of the water cycle and climate system.
190
Frequently Asked Questions
FAQ 8.1: How do land use changes affect the water cycle?
Altering land use affects the water cycle in many ways, with subsequent consequences for the whole cycle.
Urban
Precipitation
cover
Soil
moisture Runoff
Aerosols Infiltration
Water
extraction Groundwater
Surface
Vegetation
evaporation
Global Plant
warming transpiration
FAQ
FAQ 8.1, Figure 1 | Land-use changes and their consequences on the water cycle. As all the components or the water cycle are tightly connected,
changes in one aspect of the cycle affects almost all the cycle.
191
Frequently Asked Questions
FAQ 8.2 | Will Floods Become More Severe or More Frequent as a Result of Climate Change?
A warmer climate increases the amount and intensity of rainfall during wet events, and this is expected to
amplify the severity of flooding. However, the link between rainfall and flooding is complex, so while the most
severe flooding events are expected to worsen, floods could become rarer in some regions.
Floods are a natural and important part of the water cycle but they can also threaten lives and safety, disrupt
human activities, and damage infrastructure. Most inland floods occur when rivers overtop their banks (fluvial
flooding) or when intense rainfall causes water to build up and overflow locally (pluvial flooding). Flooding is
also caused by coastal inundation by the sea, rapid seasonal melting of snow, and the accumulation of debris,
such as vegetation or ice, that stops water from draining away.
Climate change is already altering the location, frequency and severity of flooding. Close to the coasts, rising
sea levels increasingly cause more frequent and severe coastal flooding, and the severity of these floods is
exacerbated when combined with heavy rainfall. The heavy and sustained rainfall events responsible for most
inland flooding are becoming more intense in many areas as the climate warms because air near Earth’s surface
can carry around 7% more water in its gas phase (vapour) for each 1°C of warming. This extra moisture is drawn
into weather systems, fueling heavier rainfall (FAQ 8.2, Figure 1).
A warming climate also affects wind patterns, how storms form and evolve, and the pathway those storms
usually travel. Warming also increases condensation rates, which in turn releases extra heat that can energize
storm systems and further intensify rainfall. On the other hand, this energy release can also inhibit the uplift
required for cloud development, while increases in particle pollution can delay rainfall but invigorate storms.
These changes mean that the character of precipitation events (how often, how long lasting and how heavy they
are) will continue to change as the climate warms.
In addition to climate change, the location, frequency and timing of the heaviest rainfall events and worst
flooding depend on natural fluctuations in wind patterns that make some regions unusually wet or dry for
months, years, or even decades. These natural variations make it difficult to determine whether heavy rainfall
events are changing locally as a result of global warming. However, when natural weather patterns bring heavy
FAQ and prolonged rainfall in a warmer climate, the intensity is increased by the larger amount of moisture in the air.
An increased intensity and frequency of record-breaking daily rainfall has been detected for much of the land
surface where good observational records exist, and this can only be explained by human-caused increases in
atmospheric greenhouse gas concentrations. Heavy rainfall is also projected to become more intense in the future
for most places. So, where unusually wet weather events or seasons occur, the rainfall amounts are expected to
be greater in the future, contributing to more severe flooding.
However, heavier rainfall does not always lead to greater flooding. This is because flooding also depends
upon the type of river basin, the surface landscape, the extent and duration of the rainfall, and how wet the
ground is before the rainfall event (FAQ 8.2, Figure 1). Some regions will experience a drying in the soil as the
climate warms, particularly in subtropical climates, which could make floods from a rainfall event less probable
because the ground can potentially soak up more of the rain. On the other hand, less frequent but more intense
downpours can lead to dry, hard ground that is less able to soak up heavy rainfall when it does occur, resulting
in more runoff into lakes, rivers and hollows. Earlier spring snowmelt combined with more precipitation falling
as rain rather than snow can trigger flood events in cold regions. Reduced winter snow cover can, in contrast,
decrease the chance of flooding arising from the combination of rainfall and rapid snowmelt. Rapid melting of
glaciers and snow in a warming climate is already increasing river flow in some regions, but as the volumes of ice
diminish, flows will peak and then decline in the future. Flooding is also affected by changes in the management
of the land and river systems. For example, clearing forests for agriculture or building cities can make rainwater
flow more rapidly into rivers or low-lying areas. On the other hand, increased extraction of water from rivers can
reduce water levels and the likelihood of flooding.
192
Frequently Asked Questions
A mix of both increases and decreases in flooding have been observed in some regions and these changes have
been attributed to multiple causes, including changes in snowmelt, soil moisture and rainfall. Although we
know that a warming climate will intensify rainfall events, local and regional trends are expected to vary in
both direction and magnitude as global warming results in multiple, and sometimes counteracting, influences.
However, even accounting for the many factors that generate flooding, when weather patterns cause flood
events in a warmer future, these floods will be more severe.
– – – – – –
+ + +
Worsening coastal
flooding from sea level Urbanisation
More moisture fuels rise and combined
heavier precipitation increases runoff
with heavier rainfall and flash flooding
+ + +
+ + –
– FAQ
Greater capacity of
Decreased glaciers
drier soils to soak up
and snow reduce
water from
river flow
sustained rainfall
+ +
+ +
– Less severe flooding
+
Heavier rainfall
+ More severe flooding increases flood severity
FAQ 8.2, Figure 1 | Schematic illustrating factors important in determining changes in heavy precipitation and flooding.
193
Frequently Asked Questions
FAQ 8.3 | What Causes Droughts, and Will Climate Change Make Them Worse?
Droughts usually begin as a deficit of precipitation, but then propagate to other parts of the water cycle (soils,
rivers, snow/ice and water reservoirs). They are also influenced by factors like temperature, vegetation and
human land and water management. In a warmer world, evaporation increases, which can make even wet
regions more susceptible to drought.
A drought is broadly defined as drier than normal conditions; that is, a moisture deficit relative to the average
water availability at a given location and season. Since they are locally defined, a drought in a wet place will
not have the same amount of water deficit as a drought in a dry region. Droughts are divided into different
categories based on where in the water cycle the moisture deficit occurs: meteorological drought (precipitation),
hydrological drought (runoff, streamflow, and reservoir storage), and agricultural or ecological drought (plant
stress from a combination of evaporation and low soil moisture). Special categories of drought also exist. For
example, a snow drought occurs when winter snowpack levels are below average, which can cause abnormally
low streamflow in subsequent seasons. And while many drought events develop slowly over months or years,
some events, called flash droughts, can intensify over the course of days or weeks. One such event occurred
in 2012 in the Midwestern region of North America and had a severe impact on agricultural production, with
losses exceeding $30 billion US dollars. Droughts typically only become a concern when they adversely affect
people (reducing water available for municipal, industrial, agricultural, or navigational needs) and/or ecosystems
(adverse effects on natural flora and fauna). When a drought lasts for a very long time (more than two decades)
it is sometimes called a megadrought.
Most droughts begin when precipitation is below normal for an extended period of time (meteorological
drought). This typically occurs when high pressure in the atmosphere sets up over a region, reducing cloud
formation and precipitation over that area and deflecting away storms. The lack of rainfall then propagates
across the water cycle to create agricultural drought in soils and hydrological drought in waterways. Other
processes act to amplify or alleviate droughts. For example, if temperatures are abnormally high, evaporation
increases, drying out soils and streams and stressing plants beyond what would have occurred from the lack of
precipitation alone. Vegetation can play a critical role because it modulates many important hydrologic processes
FAQ (soil water, evapotranspiration, runoff). Human activities can also determine how severe a drought is. For
example, irrigating croplands can reduce the socio-economic impact of a drought; at the same time, depletion of
groundwater in aquifers can make a drought worse.
The effect of climate change on drought varies across regions. In the subtropical regions like the Mediterranean,
southern Africa, south-western Australia and south-western South America, as well as tropical Central America,
western Africa and the Amazon basin, precipitation is expected to decline as the world warms, increasing the
possibility that drought will occur throughout the year (FAQ 8.3, Figure 1). Warming will decrease snowpack,
amplifying drought in regions where snowmelt is an important water resource (such as in south-western South
America). Higher temperatures lead to increased evaporation, resulting in soil drying, increased plant stress, and
impacts on agriculture, even in regions where large changes in precipitation are not expected (such as central
and northern Europe). If emissions of greenhouse gases are not curtailed, about a third of global land areas
are projected to suffer from at least moderate drought by 2100. On the other hand, some areas and seasons
(such as high-latitude regions in North America and Asia, and the South Asian monsoon region) may experience
increases in precipitation as a result of climate change, which will decrease the likelihood of droughts. FAQ 8.3,
Figure 1 highlights the regions where climate change is expected to increase the severity of droughts.
194
Frequently Asked Questions
FAQ 8.3, Figure 1 | Schematic map highlighting in brown the regions where droughts are expected to become worse as a result of climate
change. This pattern is similar regardless of the emissions scenario; however, the magnitude of change increases under higher emissions. FAQ
195
Frequently Asked Questions
FAQ 9.1 | Can Continued Melting of the Greenland and Antarctic Ice Sheets Be Reversed? How Long
Would It Take for Them to Grow Back?
Evidence from the distant past shows that some parts of the Earth system might take hundreds to thousands of
years to fully adjust to changes in climate. This means that some of the consequences of human-induced climate
change will continue for a very long time, even if atmospheric heat-trapping gas levels and global temperatures
are stabilized or reduced in the future. This is especially true for the Greenland and Antarctic ice sheets, which
grow much more slowly than they retreat. If the current melting of these ice sheets continues for long enough,
it becomes effectively irreversible on human time scales, as does the sea level rise caused by that melting.
Humans are changing the climate and there are mechanisms that amplify the warming in the polar regions
(Arctic and Antarctic). The Arctic is already warming faster than anywhere else (see FAQ 4.3). This is significant
because these colder high latitudes are home to our two remaining ice sheets: Antarctica and Greenland. Ice
sheets are huge reservoirs of frozen freshwater, built up by tens of thousands of years of snowfall. If they were
to completely melt, the water released would raise global sea level by about 65 m. Understanding how these ice
sheets are affected by warming of nearby ocean and atmosphere is therefore critically important. The Greenland
and Antarctic ice sheets are already slowly responding to recent changes in climate, but it takes a long time
for these huge masses of ice to adjust to changes in global temperature. That means that the full effects of
a warming climate may take hundreds or thousands of years to play out. An important question is whether these
changes can eventually be reversed, once levels of greenhouse gases in the atmosphere are stabilized or reduced
by humans and natural processes. Records from the past can help us answer this question.
For at least the last 800,000 years, the Earth has followed cycles of gradual cooling followed by rapid warming
caused by natural processes. During cooling phases, more and more ocean water is gradually deposited as
snowfall, causing ice sheets to grow and sea level to slowly decrease. During warming phases, the ice sheets
melt more quickly, resulting in more rapid rises in sea level (FAQ 9.1, Figure 1). Ice sheets build up very slowly
because growth relies on the steady accumulation of falling snow that eventually compacts into ice. As the
climate cools, areas that can accumulate snow expand, reflecting back more sunlight that otherwise would keep
the Earth warmer. This means that, once started, glacial climates develop rapidly. However, as the climate cools,
FAQ
the amount of moisture that the air can hold tends to decrease. As a result, even though glaciations begin quite
quickly, it takes tens of thousands of years for ice sheets to grow to a point where they are in balance with the
colder climate.
Ice sheets retreat more quickly than they grow because of processes that, once triggered, drive self-reinforcing
ice loss. For ice sheets that are mostly resting on bedrock above sea level – like the Greenland Ice Sheet – the main
self-reinforcing loop that affects them is the ‘elevation–mass balance feedback’ (FAQ 9.1, Figure 1, right). In this
situation, the altitude of the ice-sheet surface decreases as it melts, exposing the sheet to warmer air. The lowered
surface then melts even more, lowering it faster still, until eventually the whole ice sheet disappears. In places
where the ice sheet rests instead on bedrock that is below sea level, and which also deepens inland, including
many parts of the Antarctic Ice Sheet, an important process called ‘marine ice sheet instability’ is thought to drive
rapid retreat (FAQ 9.1, Figure 1, left). This happens when the part of the ice sheet that is surrounded by sea water
melts. That leads to additional thinning, which in turn accelerates the motion of the glaciers that feed into these
areas. As the ice sheet flows more quickly into the ocean, more melting takes place, leading to more thinning
and even faster flow that brings ever-more glacier ice into the ocean, ultimately driving rapid deglaciation of
whole ice-sheet drainage basins.
These (and other) self-reinforcing processes explain why relatively small increases in temperature in the past led
to very substantial sea level rise over centuries to millennia, compared to the many tens of thousands of years it
takes to grow the ice sheets that lowered the sea level in the first place. These insights from the past imply that,
if human-induced changes to the Greenland and Antarctic ice sheets continue for the rest of this century, it will
take thousands of years to reverse that melting, even if global air temperatures decrease within this or the next
century. In this sense, these changes are therefore irreversible, since the ice sheets would take much longer to
regrow than the decades or centuries for which modern society is able to plan.
196
Frequently Asked Questions
10
0
Slow ice sheet growth
Global mean sea level (m)
-20
Rapid ice sheet retreat
-40
-60
-80
-100
-120
-140
140,000 120,000 100,000 80,000 60,000 40,000 20,000 Today
Years before present
Ice sheet
Ocean
Bedrock Bedrock Ice sheet
When bedrock dips seaward or is flat, the retreat The ice sheet is very thick therefore its surface is
stops when warming stops. When ice sheet retreats, very high and the air at high altitude is very cold
less ice is released into ocean
When bedrock dips landward the retreat is quick and As the ice sheet melts, its surface goes down
self-sustained. When ice sheet retreats, more ice until it reaches a threshold, where the surrounding
is released into ocean – ice sheet retreats further air is warmer and melts the ice even more quickly
FAQ 9.1, Figure 1 | Ice sheets growth and decay. (Top) Changes in ice-sheet volume modulate sea level variations. The grey line depicts data from a range
of physical environmental sea level recorders such as coral reefs while the blue line is a smoothed version of it. (Bottom left) Example of destabilization
mechanism in Antarctica. (Bottom right) Example of destabilization mechanism in Greenland.
197
Frequently Asked Questions
FAQ 9.2 | How Much Will Sea Level Rise in the Next Few Decades?
As of 2018, global average sea level was about 15–25 cm higher than in 1900, and 7–15 cm higher than in 1971.
Sea level will continue to rise by an additional 10–25 cm by 2050. The major reasons for this ongoing rise in sea
level are the thermal expansion of seawater as its temperature increases, and the melting of glaciers and ice
sheets. Local sea level changes can be larger or smaller than the global average, with the smallest changes in
formerly glaciated areas, and the largest changes in low-lying river delta regions.
Across the globe, sea level is rising, and the rate of increase has accelerated. Sea level increased by about 4 mm
per year from 2006 to 2018, which was more than double the average rate over the 20th century. Rise during
the early 1900s was due to natural factors, such as glaciers catching up to warming that occurred in the Northern
Hemisphere during the 1800s. However, since at least 1970, human activities have been the dominant cause of
global average sea level rise, and they will continue to be for centuries into the future.
Sea level rises either through warming of ocean waters or the addition of water from melting ice and bodies of
water on land. Expansion due to warming caused about 50% of the rise observed from 1971 to 2018. Melting
glaciers contributed about 22% over the same period. Melting of the two large ice sheets in Greenland and
Antarctica has contributed about 13% and 7%, respectively, during 1971 to 2018, but melting has accelerated
in the recent decades, increasing their contribution to 22% and 14% since 2016. Another source is changes in
land-water storage: reservoirs and aquifers on land have reduced, which contributed about an 8% increase
in sea level.
By 2050, sea level is expected to rise an additional 10–25 cm whether or not greenhouse gas emissions are reduced
(FAQ 9.2, Figure 1). Beyond 2050, the amount by which sea level will rise is more uncertain. The accumulated
total emissions of greenhouse gases over the upcoming decades will play a big role beyond 2050, especially in
determining where sea level rise and ice-sheet changes eventually level off.
Even if net zero emissions are reached, sea level rise will continue because the deep ocean will continue to warm
and ice sheets will take time to catch up to the warming caused by past and present emissions: ocean and ice
sheets are slow to respond to environmental changes (see FAQ 5.3). Some projections under low emissions show
FAQ sea level rise continuing as net zero is approached at a rate comparable to today (3–8 mm per year by 2100 versus
3–4 mm per year in 2015), while others show substantial acceleration to more than five times the present rate by
2100, especially if emissions continue to be high and processes that accelerate retreat of the Antarctic Ice Sheet
occur widely (FAQ 9.1).
Sea level rise will increase the frequency and severity of extreme sea level events at coasts (see FAQ 8.2), such as
storm surges, wave inundation and tidal floods: risk can be increased by even small changes in global average
sea level. Scientists project that, in some regions, extreme sea level events that were recently expected once in
100 years will occur annually at 20–25% of locations by 2050 regardless of emissions, but by 2100 emissions choice
will matter: annually at 60% of locations for low emissions, and at 80% of locations under strong emissions.
In many places, local sea level change will be larger or smaller than the global average. From year to year and
place to place, changes in ocean circulation and wind can lead to local sea level change. In regions where large
ice sheets, such as the Fennoscandian in Eurasia and the Laurentide and Cordilleran in North America, covered
the land during the last ice age, the land is still slowly rising up now that the extra weight of the ice sheets is
gone. This local recovery is compensating for global sea level rise in these regions and can even lead to local
decrease in sea level. In regions just beyond where the former ice sheets reached and the Earth bulged upwards,
the land is now falling and, as a result, local sea level rise is faster than the global rate. In many regions within
low-lying delta regions (such as New Orleans and the Ganges–Brahmaputra delta), the land is rapidly subsiding
(sinking) because of human activities such as building dams or groundwater and fossil fuel extraction. Further,
when an ice sheet melts, it has less gravitational pull on the ocean water nearby. This reduction in gravitational
attraction causes sea level to fall close to the (now less-massive) ice sheet while causing sea level to rise farther
away. Melt from a polar ice sheet therefore raises sea level most in the opposite hemisphere or in low latitudes –
amounting to tens of centimetres difference in rise between regions by 2100.
198
Frequently Asked Questions
FAQ 9.2: How much will sea level rise in the next few decades?
Emissions scenarios influence little sea level rise of the coming decades but has a huge effect
on sea level at the end of the century.
In 2018 In 2050 In 2100
Contributions
cm of sea level rise
(since 1971)
Observations
Antarctic ice sheet
Greenland ice sheet Low-emission scenario – SSP1-2.6
Glaciers
Ocean thermal expansion
Water stored on land
50
SSP1-2.6
FAQ 9.2, Figure 1 | Observed and projected global mean sea level rise and the contributions from its major constituents.
199
Frequently Asked Questions
The Gulf Stream is part of two circulation patterns in the North Atlantic: the Atlantic Meridional Overturning
Circulation (AMOC) and the North Atlantic subtropical gyre. Based on models and theory, scientific studies
indicate that, while the AMOC is expected to slow in a warming climate, the Gulf Stream will not change much
and would not shut down totally, even if the AMOC did. Most climate models project that the AMOC slows in the
later 21st century under most emissions scenarios, with some models showing it slowing even sooner. The Gulf
Stream affects the weather and sea level, so if it slows, North America will see higher sea levels and Europe’s
weather and rate of relative warming will be affected.
The Gulf Stream is the biggest current in the North Atlantic Ocean. It transports about 30 billion kilograms
of water per second northward past points on the east coast of North America. It is a warm current, with
temperatures 5°C to 15°C warmer than surrounding waters, so it carries warmer water (thermal energy) from its
southern origins and releases warmth to the atmosphere and surrounding water.
The Gulf Stream is part of two major circulation patterns, the Atlantic Meridional Overturning Circulation (AMOC)
and the North Atlantic Subtropical Gyre (FAQ 9.3, Figure 1). The rotation of the Earth causes the big currents in
both circulations to stay on the western side of their basin, which in the Atlantic means the circulations combine
to form the Gulf Stream. Other large currents contribute to gyres, such as the Kuroshio in the North Pacific
and the East Australian Current in the South Pacific, but the Gulf Stream is special in its dual role. There is no
comparable deep overturning circulation in the North Pacific to the AMOC, so the Kuroshio plays only one role
as part of a gyre.
The gyres circulate surface waters and result primarily from winds driving the circulation. These winds are not
expected to change much and so neither will the gyres, which means the gyre portion of the Gulf Stream and the
Kuroshio will continue to transport thermal energy poleward from the equator much as they do now. The gyre
contribution to the Gulf Stream is 2 to 10 times larger than the AMOC contribution.
The Gulf Stream’s role in the AMOC is supplying surface source water that cools, becomes denser and sinks to
form cold, deep waters that travel back equatorward, spilling over features on the ocean floor and mixing with
FAQ
other deep Atlantic waters to form a southward current at a depth of about 1500 metres beneath the Gulf
Stream. This overturning flow is the AMOC, with the Gulf Stream in the upper kilometre flowing northward,
and the colder deep water flowing southward.
The AMOC is expected to slow over the coming centuries. One reason why is freshening of the ocean waters:
by meltwater from Greenland, changing Arctic sea ice, and increased precipitation over warmer northern seas.
An array of moorings across the Atlantic has been monitoring the AMOC since 2004, with recently expanded
capabilities. The monitoring of the AMOC has not been long enough for a trend to emerge from variability
and detect long-term changes that may be underway (see FAQ 1.2). Other indirect signs may indicate slowing
overturning – for example, slower warming where the Gulf Stream’s surface waters sink. Climate models show
that this ‘cold spot’ of slower-than-average warming occurs as the AMOC weakens, and they project that this
will continue. Paleoclimate evidence indicates that the AMOC changed significantly in the past, especially during
transitions from colder climates to warmer ones, but that it has been stable for 8000 years.
What happens if the AMOC slows in a warming world? The atmosphere adjusts somewhat by carrying more heat,
compensating partly for the decreases in heat carried by AMOC. But the ‘cold spot’ makes parts of Europe warm
more slowly. Models indicate that weather patterns in Greenland and around the Atlantic will be affected, with
reduced precipitation in the mid-latitudes, changing strong precipitation patterns in the tropics and Europe, and
stronger storms in the North Atlantic storm track. The slowing of this current combined with the rotation of the
Earth means that sea level along North America rises as the AMOC contribution to the Gulf Stream slows.
The North Atlantic is not the only site of sensitive meridional overturning. Around Antarctica, the world’s densest
seawater is formed by freezing into sea ice, leaving behind salty, cold water that sinks to the bottom and spreads
northward. Recent studies show that melting of the Antarctic Ice Sheet and changing winds over the Southern
Ocean can affect this southern meridional overturning, affecting regional weather.
200
Frequently Asked Questions
1
Warm surface current 2 Much less
2
driven by winds and heat and water
replenishing sinking Water and
heat transferred are transferred
from the tropics to
GULF northern latitudes
GULF
STREAM STREAM
GYRE GYRE
3
The Gulf Stream
weakens but the
4 portion pushed by
winds remains
The cold deep water
is exported southward
FAQ 9.3, Figure 1 | Horizontal (gyre) and vertical (Atlantic Meridional Overturning Circulation, AMOC) circulations in the Atlantic today FAQ
(left) and in a warmer world (right). The Gulf Stream is a warm current composed of both circulations.
201
Frequently Asked Questions
FAQ 10.1 | How Can We Provide Useful Climate Information for Regional Stakeholders?
The world is physically and culturally diverse, and the challenges posed by climate change vary by region and
location. Because climate change affects so many aspects of people’s daily work and living, climate change
information can help with decision-making, but only when the information is relevant for the people involved in
making those decisions. Users of climate information may be highly diverse, ranging from professionals in areas
such as human health, agriculture or water management to a broader community that experiences the impacts
of changing climate. Providing information that supports response actions thus requires engaging all relevant
stakeholders, their knowledge and their experiences, formulating appropriate information, and developing
a mutual understanding of the usefulness and limitations of the information.
The development, delivery, and use of climate change information requires engaging all parties involved:
those producing the climate data and related knowledge, those communicating it, and those who combine
that information with their knowledge of the community, region or activity that climate change may impact.
To be successful, these parties need to work together to explore the climate data and thus co-develop the
climate information needed to make decisions or solve problems, distilling output from the various sources
of climate knowledge into relevant climate information. Effective partnerships recognize and respond to the
diversity of all parties involved (including their values, beliefs and interests), especially when they involve
culturally diverse communities and their indigenous and local knowledge of weather, climate and their society.
This is particularly true for climate change – a global issue posing challenges that vary by region. By recognizing
this diversity, climate information can be relevant and credible, most notably when conveying the complexity of
risks for human systems and ecosystems and for building resilience.
Constructing useful climate information requires considering all available sources in order to capture the fullest
possible representation of projected changes and distilling the information in a way that meets the needs of the
stakeholders and communities impacted by the changes. For example, climate scientists can provide information
on future changes by using simulations of global and/or regional climate and inferring changes in the weather
behaviour influencing a region. An effective distillation process (FAQ 10.1, Figure 1) engages with the intended
recipients of the information, especially stakeholders whose work involves non-climatic factors, such as human
FAQ health, agriculture or water resources. The distillation evaluates the accuracy of all information sources
(observations, simulations, expert judgement), weighs the credibility of possible conflicting information, and
arrives at climate information that includes estimating the confidence a user should have in it. Producers of
climate data should further recognize that the geographic regions and time periods governing stakeholders’
interest (for example, the growing season of an agricultural zone) may not align well with the time and space
resolution of available climate data; thus additional model development or data processing may be required to
extract useful climate information.
One way to distil complex information for stakeholder applications is to connect this information to experiences
stakeholders have already had through storylines as plausible unfoldings of weather and climate events related
to stakeholders’ experiences. Dialogue between stakeholders and climate scientists can determine the most
relevant experiences to evaluate for possible future behaviour. The development of storylines uses the experience
and expertise of stakeholders, such as water-resource managers and health professionals, who seek to develop
appropriate response measures. Storylines are thus a pathway through the distillation process that can make
climate information more accessible and physically comprehensible. For example, a storyline may take a common
experience like an extended drought, with depleted water availability and damaged crops, and show how
droughts may change in the future, perhaps with even greater precipitation deficits or longer duration. With
appropriate choices, storylines can engage nuances of the climate information in a meaningful way by building
on common experiences, thus enhancing the information’s usefulness.
Forging partnerships among all involved with producing, exploring and distilling climate data into climate
information is at the centre of creating stakeholder-relevant information. These partnerships can occur through
direct interaction between climate scientists and stakeholders as well as through organizations that have
emerged to facilitate this process, such as climate services, national and regional climate forums, and consulting
firms providing specialized climate information. These so-called ‘boundary organizations’ can serve the varied
needs of all who would fold climate information into their decision processes. All of these partnerships are vital
202
Frequently Asked Questions
for arriving at climate information that responds to physical and cultural diversity and to challenges posed by
climate change that can vary region-by-region around the world.
FAQ 10.1: How can scientists provide useful regional climate information?
In decision-making, climate information is more useful if the physical and cultural diversity across the world is considered.
Distillation
involving
scientists and
stakeholders
FAQ
Human health Agriculture Water resources Transportation Fisheries Natural Human Energy
management ecosystems habitations resources
FAQ 10.1, Figure 1 | Climate information for decision makers is more useful if the physical and cultural diversity across the world is
considered. The figure illustrates schematically the broad range of knowledge that must be blended with the diversity of users to distil information that will
have relevance and credibility. This blending or distillation should engage the values and knowledge of both the stakeholders and the scientists. The bottom
row contains examples of stakeholders’ interests and is not all-inclusive. As part of the distillation, the outcomes can advance the United Nations’ Sustainable
Development Goals, covered in part by these examples.
203
Frequently Asked Questions
Urban areas experience air temperatures that can be several degrees Celsius warmer than surrounding areas,
especially during the night. This ‘urban heat island’ effect results from several factors, including reduced
ventilation and heat trapping due to the close proximity of tall buildings, heat generated directly from human
activities, the heat-absorbing properties of concrete and other urban building materials, and the limited amount
of vegetation. Continuing urbanization and increasingly severe heatwaves under climate change will further
amplify this effect in the future.
Today, cities are home to 55% of the world’s population. This number is increasing, and every year cities welcome
67 million new residents, 90% of whom are moving to cities in developing countries. By 2030, almost 60% of
the world’s population is expected to live in urban areas. Cities and their inhabitants are highly vulnerable to
weather and climate extremes, particularly heatwaves, because urban areas already are local hotspots. Cities are
generally warmer – up to several degrees Celsius at night – than their surroundings. This warming effect, called
the urban heat island, occurs because cities both receive and retain more heat than the surrounding countryside
areas and because natural cooling processes are weakened in cities compared to rural areas.
Three main factors contribute to amplify the warming of urban areas (orange bars in FAQ 10.2, Figure 1). The
strongest contribution comes from urban geometry, which depends on the number of buildings, their size and
their proximity. Tall buildings close to each other absorb and store heat and also reduce natural ventilation.
Human activities, which are very concentrated in cities, also directly warm the atmosphere locally, due to heat
released from domestic and industrial heating or cooling systems, running engines, and other sources. Finally,
urban warming also results directly from the heat-retaining properties of the materials that make up cities,
including concrete buildings, asphalt roadways, and dark rooftops. These materials are very good at absorbing
and retaining heat, and then re-emitting that heat at night.
The urban heat island effect is further amplified in cities that lack vegetation and water bodies, both of which can
strongly contribute to local cooling (green bars in FAQ 10.2, Figure 1). This means that when enough vegetation
and water are included in the urban fabric, they can counterbalance the urban heat island effect, to the point of
FAQ even cancelling out the urban heat island effect in some neighbourhoods.
The urban heat island phenomenon is well-known and understood. For instance, temperature measurements
from thermometers located in cities are corrected for this effect when global warming trends are calculated.
Nevertheless, observations, including long-term measurements of the urban heat island effect are currently too
limited to allow a full understanding of how the urban heat island varies across the world and across different
types of cities and climatic zones, or how this effect will evolve in the future.
As a result, it is hard to assess how climate change will affect the urban heat island effect, and various studies
disagree. Two things are, however, very clear. First, future urbanization will expand the urban heat island areas,
thereby amplifying future warming in many places all over the world. In some places, the nighttime warming
from the urban heat island effect could even be on the same order of magnitude as the warming expected from
human-induced climate change. Second, more intense, longer and more frequent heatwaves caused by climate
change will more strongly impact cities and their inhabitants, because the extra warming from the urban heat
island effect will exacerbate the impacts of climate change.
In summary, cities are currently local hotspots because their structure, material and activities trap and release
heat and reduce natural cooling processes. In the future, climate change will, on average, have a limited effect
on the magnitude of the urban heat island itself, but ongoing urbanization together with more frequent, longer
and warmer heatwaves will make cities more exposed to global warming.
204
Frequently Asked Questions
Heat from
human activities
Domestic/
industrial heating
Heat-retaining
properties
Buildings and
road materials
Variations across
different climates
Water
Sea, river,
lakes, irrigation
Cities often lack
vegetation and water
Vegetation
Parks, forests,
gardens
FAQ
-3 -2 -1 0 +1 +2
COOLING Local effect on temperature (ºC) WARMING
FAQ 10.2, Figure 1 | Efficiency of the various factors at warming up or cooling down neighbourhoods of urban areas. Overall, cities tend to be
warmer than their surroundings. This is called the ‘urban heat island’ effect. The hatched areas on the bars show how the strength of the warming or cooling
effects of each factor varies depending on the local climate. For example, vegetation has a stronger cooling effect in temperate and warm climates. Further
details on data sources are available in the chapter data table (Table 10.SM.11).
205
Frequently Asked Questions
Human-caused climate change alters the frequency and intensity of climate variables (e.g., surface temperature)
and phenomena (e.g., tropical cyclones) in a variety of ways. We now know that the ways in which average and
extreme conditions have changed (and will continue to change) depend on the variable and the phenomenon
being considered. Changes in local surface temperature extremes closely follow the corresponding changes in
local average surface temperatures. On the contrary, changes in precipitation extremes (heavy precipitation)
generally do not follow those in average precipitation, and can even move in the opposite direction (e.g., with
average precipitation decreasing but extreme precipitation increasing).
Climate change will manifest very differently depending on which region, season and variable we are interested
in. For example, over some parts of the Arctic, temperatures will warm at rates about three to four times higher
during winter compared to summer months. And in summer, most of northern Europe will experience larger
temperatures increases than most places in south-east South America and Australasia, with differences that can
be larger than 1°C, depending on the level of global warming. In general, differences across regions and seasons
arise because the underlying physical processes differ drastically across regions and seasons.
Climate change will also manifest differently for different weather regimes and can lead to contrasting changes
in average and extreme conditions. Observations of the recent past and climate model projections show that,
in most places, changes in daily temperatures are dominated by a general warming where the climatological
average and extreme values are shifted towards higher temperatures, making warm extremes more frequent and
cold extremes less frequent. The top panels in FAQ 11.1, Figure 1 show projected changes in surface temperature
for long-term average conditions (left) and for extreme hot days (right) during the warm season (summer in mid-
to high latitudes). Projected increases in long-term average temperature differ substantially between different
places, varying from less than 3°C in some places in central South Asia and southern South America to over 7°C
in some places in North America, North Africa and the Middle East. Changes in extreme hot days follow changes
in average conditions quite closely, although, in some places, the warming rates for extremes can be intensified
(e.g., southern Europe and the Amazon basin) or weakened (e.g., northern Asia and Greenland) compared to
FAQ average values.
Recent observations and global and regional climate model projections point to changes in precipitation
extremes (including both rainfall and snowfall extremes) differing drastically from those in average precipitation.
The bottom panels in FAQ 11.1, Figure 1 show projected changes in the long-term average precipitation (left)
and in heavy precipitation (right). Averaged precipitation changes show striking regional differences, with
substantial drying in places such as southern Europe and northern South America and wetting in places such as
the Middle East and southern South America. Changes in extreme precipitation are much more uniform, with
systematic increases over nearly all land regions. The physical reasons behind the different responses of averaged
and extreme precipitation are now well understood. The intensification of extreme precipitation is driven by
the increase in atmospheric water vapour (about 7% per 1°C of warming near the surface), although this is
modulated by various dynamical changes. In contrast, changes in average precipitation are driven not only by
moisture increases but also by slower processes that constrain future changes over the globe to only 2–3% per
1°C of warming near the surface.
In summary, the specific relationship between changes in average and extreme conditions strongly depends on
the variable or phenomenon being considered. At the local scale, average and extreme surface temperature
changes are strongly related, while average and extreme precipitation changes are often weakly related. For
both variables, the changes in average and extreme conditions vary strongly across different places due to the
effect of local and regional processes.
206
Frequently Asked Questions
FAQ 11.1: How will changes in climate extremes compare with changes in climate averages?
The direction and magnitude of future changes in climate extremes and averages depend on the variable considered.
Future changes
in temperature
averages and
extremes will be
similar
0°C 2 4 6 8°C
Warmer
Climate average Climate extreme
Future changes
in precipitation
averages and -
extremes can be
very different
FAQ 11.1, Figure 1 | Global maps of future changes in surface temperature (top panels) and precipitation (bottom panels) for long-term FAQ
average (left) and extreme conditions (right). All changes were estimated using the Coupled Model Intercomparison Project Phase 6 (CMIP6) ensemble
median for a scenario with a global warming of 4°C relative to 1850–1900 temperatures. Average surface temperatures refer to the warmest three-month
season (summer in mid- to high latitudes) and extreme temperatures refer to the hottest day in a year. Precipitation changes, which can include both rainfall
and snowfall changes, are normalized by 1850–1900 values and shown as a percentage; extreme precipitation refers to the largest daily precipitation in a year.
207
Frequently Asked Questions
FAQ 11.2 | Will Unprecedented Extremes Occur As a Result Of Human-Induced Climate Change?
Climate change has already increased the magnitude and frequency of extreme hot events and decreased
the magnitude and frequency of extreme cold events, and, in some regions, intensified extreme precipitation
events. As the climate moves away from its past and current states, we will experience extreme events that
are unprecedented, either in magnitude, frequency, timing or location. The frequency of these unprecedented
extreme events will rise with increasing global warming. Additionally, the combined occurrence of multiple
unprecedented extremes may result in large and unprecedented impacts.
Human-induced climate change has already affected many aspects of the climate system. In addition to the increase
in global surface temperature, many types of weather and climate extremes have changed. In most regions, the
frequency and intensity of hot extremes have increased and those of cold extremes have decreased. The frequency
and intensity of heavy precipitation events have increased at the global scale and over a majority of land regions.
Although extreme events such as land and marine heatwaves, heavy precipitation, drought, tropical cyclones, and
associated wildfires and coastal flooding have occurred in the past and will continue to occur in the future, they
often come with different magnitudes or frequencies in a warmer world. For example, future heatwaves will last
longer and have higher temperatures, and future extreme precipitation events will be more intense in several
regions. Certain extremes, such as extreme cold, will be less intense and less frequent with increasing warming.
Unprecedented extremes – that is, events not experienced in the past – will occur in the future in five different ways
(FAQ 11.2, Figure 1). First, events that are considered to be extreme in the current climate will occur in the future
with unprecedented magnitudes. Second, future extreme events will also occur with unprecedented frequency.
Third, certain types of extremes may occur in regions that have not previously encountered those types of events.
For example, as the sea level rises, coastal flooding may occur in new locations, and wildfires are already occurring
in areas, such as parts of the Arctic, where the probability of such events was previously low. Fourth, extreme events
may also be unprecedented in their timing. For example, extremely hot temperatures may occur either earlier or
later in the year than they have in the past.
208
Frequently Asked Questions
FAQ 11.3 | Did Climate Change Cause That Recent Extreme Event In My Country?
While it is difficult to identify the exact causes of a particular extreme event, the relatively new science of event
attribution is able to quantify the role of climate change in altering the probability and magnitude of some
types of weather and climate extremes. There is strong evidence that characteristics of many individual extreme
events have already changed because of human-driven changes to the climate system. Some types of highly
impactful extreme weather events have occurred more often and have become more severe due to these human
influences. As the climate continues to warm, the observed changes in the probability and/or magnitude of some
extreme weather events will continue as the human influences on these events increase.
It is common to question whether human-caused climate change caused a major weather- and climate-related
disaster. When extreme weather and climate events do occur, both exposure and vulnerability play an important
role in determining the magnitude and impacts of the resulting disaster. As such, it is difficult to attribute a specific
disaster directly to climate change. However, the relatively new science of event attribution enables scientists to
attribute aspects of specific extreme weather and climate events to certain causes. Scientists cannot answer directly
whether a particular event was caused by climate change, as extremes do occur naturally, and any specific weather
and climate event is the result of a complex mix of human and natural factors. Instead, scientists quantify the relative
importance of human and natural influences on the magnitude and/or probability of specific extreme weather
events. Such information is important for disaster risk reduction planning, because improved knowledge about
changes in the probability and magnitude of relevant extreme events enables better quantification of disaster risks.
On a case-by-case basis, scientists can now quantify the contribution of human influences to the magnitude and
probability of many extreme events. This is done by estimating and comparing the probability or magnitude of
the same type of event between the current climate – including the increases in greenhouse gas concentrations
and other human influences – and an alternate world
where the atmospheric greenhouse gases remained
FAQ 11.3: Climate change and extreme events at pre-industrial levels. FAQ 11.3 Figure 1 illustrates
Extreme events have become more probable and more intense. this approach using differences in temperature and
Many of these changes can be attributed to human influence
on the climate. probability between the two scenarios as an example.
Both the pre-industrial (blue) and current (red) FAQ
More
probable
Current climate climates experience hot extremes, but with different
Same temperature probabilities and magnitudes. Hot extremes of a given
but more probable
temperature have a higher probability of occurrence
Same probability in the warmer current climate than in the cooler pre-
Pre-industrial but hotter industrial climate. Additionally, an extreme hot event
climate
Annual of a particular probability will be warmer in the current
chance
of event climate than in the pre-industrial climate. Climate model
simulations are often used to estimate the occurrence
of a specific event in both climates. The change in the
magnitude and/or probability of the extreme event
in the current climate compared to the pre-industrial
Less Probabilities are lower
probable for hotter (more extreme) events climate is attributed to the difference between the two
Cooler Temperature Warmer scenarios, which is the human influence.
209
Frequently Asked Questions
probability and magnitude are expected and, as a result, it will be possible to attribute future temperature and
precipitation extremes in many locations to human influences. Attributable changes may emerge for other types
of extremes as the warming signal increases.
In conclusion, human-caused global warming has resulted in changes in a wide variety of recent extreme weather
events. Strong increases in probability and magnitude, attributable to human influence, have been found for
many heatwaves and hot extremes around the world.
FAQ
210
Frequently Asked Questions
A climatic impact-driver is a physical climate condition that directly affects society or ecosystems. Climatic impact-
drivers may represent a long-term average condition (such as the average winter temperatures that affect indoor
heating requirements), a common event (such as a frost that kills off warm-season plants), or an extreme event
(such as a coastal flood that destroys homes). A single climatic impact-driver may lead to detrimental effects
for one part of society while benefiting another, while others are not affected at all. A climatic impact-driver
(or its change caused by climate change) is therefore not universally hazardous or beneficial, but we refer to it
as a ‘hazard’ when experts determine it is detrimental to a specific system.
Climate change can alter many aspects of the climate system, but efforts to identify impacts and risks usually focus
on a smaller set of changes known to affect, or potentially affect, things that society cares about. These climatic
impact-drivers (CIDs) are formally defined in this Report as ‘physical climate system conditions (e.g., means,
events, extremes) that affect an element of society or ecosystems. Depending on system tolerance, CIDs and
their changes can be detrimental, beneficial, neutral, or a mixture of each across interacting system elements
and regions’. Because people, infrastructure and ecosystems interact directly with their immediate environment,
climate experts assess CIDs locally and regionally. CIDs may relate to temperature, the water cycle, wind and
storms, snow and ice, oceanic and coastal processes or the chemistry and energy balance of the climate system.
Future impacts and risk may also be directly affected by factors unrelated to the climate (such as socio-economic
development, population growth, or a viral outbreak) that may also alter the vulnerability or exposure of systems.
CIDs capture important characteristics of the average climate and both common and extreme events that shape
society and nature (see FAQ 12.2). Some CIDs focus on aspects of the average climate (such as the seasonal
progression of temperature and precipitation, average winds or the chemistry of the ocean) that determine,
for example, species distribution, farming systems, the location of tourist resorts, the availability of water
resources and the expected heating and cooling needs for buildings in an average year. CIDs also include
common episodic events that are particularly important to systems, such as thaw events that can trigger the
development of plants in spring, cold spells that are important for fruit crop chill requirements, or frost events
that eliminate summer vegetation as winter sets in. Finally, CIDs include many extreme events connected to
impacts such as hailstorms that damage vehicles, coastal floods that destroy shoreline property, tornadoes that
FAQ
damage infrastructure, droughts that increase competition for water resources, and heatwaves that can strain
the health of outdoor labourers.
Many aspects of our daily lives, businesses and natural systems depend on weather and climate, and there is
great interest in anticipating the impacts of climate change on the things we care about. To meet these needs,
scientists engage with companies and authorities to provide climate services – meaningful and possibly actionable
climate information designed to assist decision-making. Climate science and services can focus on CIDs that
substantially disrupt systems to support broader risk management approaches. A single CID change can have
dramatically different implications for different sectors or even elements of the same sector, so engagement
between climate scientists and stakeholders is important to contextualize the climate changes that will come.
Climate services responding to planning and optimization of an activity can focus on more gradual changes in
operating climate conditions.
FAQ 12.1, Figure 1 tracks example outcomes of seasonal snow cover changes that connect climate science to the
need for mitigation, adaptation and regional risk management. The length of the season with snow on the ground
is just one of many regional climate conditions that may change in the future, and it becomes a CID because
there are many elements of society and ecosystems that rely on an expected seasonality of snow cover. Climate
scientists and climate service providers examining human-driven climate change may identify different regions
where the length of the season with snow cover could increase, decrease, or stay relatively unaffected. In each
region, change in seasonal snow cover may affect different systems in beneficial or detrimental ways (in the
latter case, changing seasonal snow cover would be a ‘hazard’), although systems such as coastal aquaculture
remain relatively unaffected. The changing profile of benefits and hazards connected to these changes in the
seasonal snow cover CID affects the profile of impacts, risks and benefits that stakeholders in the region will
grapple with in response to climate change.
211
Frequently Asked Questions
Increase
Seasonal No
change Inconsequential for coastal aquaculture
snow cover
! Impacts on snow-camouflaged species
Decrease
Beneficial for some predators
FAQ 12.1, Figure 1 | A single climatic impact-driver can affect ecosystems and society in different ways. A variety of impacts from the same
climatic impact-driver change, illustrated with the example of regional seasonal snow cover.
FAQ
212
Frequently Asked Questions
FAQ 12.2 | What Are Climatic Thresholds and Why Are They Important?
Climatic thresholds tell us about the tolerance of society and ecosystems so that we can better scrutinize the types
of climate changes that are expected to impact things we care about. Many systems have natural or structural
thresholds. If conditions exceed those thresholds, the result can be sudden changes or even collapses in health,
productivity, utility or behaviour. Adaptation and risk management efforts can change these thresholds, altering
the profile of climate conditions that would be problematic and increasing overall system resilience.
Decision makers have long observed that certain weather and climate conditions can be problematic, or
hazardous, for things they care about (i.e., things with socio-economic, cultural or intrinsic value). Many elements
of society and ecosystems operate in a suitable climate zone selected naturally or by stakeholders considering
the expected climate conditions. However, as climate change moves conditions beyond expected ranges, they
may cross a climatic ‘threshold’ – a level beyond which there are either gradual changes in system behaviour or
abrupt, non-linear and potentially irreversible impacts.
Climatic thresholds can be associated with either natural or structural tolerance levels. Natural thresholds, for
instance, include heat and humidity conditions above which humans cannot regulate their internal temperatures
through sweat, drought durations that heighten competition between species, and winter temperatures that
are lethal for pests or disease-carrying vector species. Structural thresholds include engineered limits of drainage
systems, extreme wind speeds that limit wind turbine operation, the height of coastal protection infrastructure,
and the locations of irrigation infrastructure or tropical cyclone sheltering facilities.
Thresholds may be defined according to raw values (such as maximum temperature exceeding 35°C) or percentiles
(such as the local 99th percentile daily rainfall total). They also often have strong seasonal dependence
(see FAQ 12.3). For example, the amount of snowfall that a deciduous tree can withstand depends on whether
the snowfall occurs before or after the tree sheds its leaves. Most systems respond to changes in complex ways,
and those responses are not determined solely or precisely by specific thresholds of a single climate variable.
Nonetheless, thresholds can be useful indicators of system behaviours, and an understanding of these thresholds
can help inform risk management decisions.
FAQ 12.2 Figure 1 illustrates how threshold conditions can help us understand climate conditions that are
FAQ
suitable for normal system operation and the thresholds beyond which impacts occur. Crops tend to grow most
optimally within a suitable range of daily temperatures that is influenced by the varieties being cultivated and
the way the farm is managed. As daily temperatures rise above a ‘critical’ temperature threshold, plants begin
to experience heat stress that reduces growth and may lower resulting yields. If temperatures reach a higher
‘limiting’ temperature threshold, crops may suffer leaf loss, pollen sterility, or tissue damage that can lead to
crop failure. Farmers typically select a cropping system with some consideration to the probability of extreme
temperature events that may occur within a typical season, and so identifying hot temperature thresholds helps
farmers select their seed and field management strategies as part of their overall risk management. Climate
experts may therefore aim to assist farm planning by providing information about the climate change-induced
shifts to the expected frequency of daily heat extremes that exceed crop tolerance thresholds.
Adaptation and other changes in societies and environment can shift climatic thresholds by modifying vulnerability
and exposure. For example, adaptation efforts may include breeding new crops with higher heat tolerance levels
so that corresponding dangerous thresholds occur less frequently. Likewise, increasing the height of a flood
embankment protecting a given community can increase the level of river flow that may be tolerated without
flooding, reducing the frequency of damaging floods. Stakeholders therefore benefit from climate services that
are based on a co-development process, with scientists identifying system-relevant thresholds and developing
tailored climatic impact-driver indices that represent these thresholds (FAQ 12.1). These thresholds help focus the
provision of action-relevant climate information for adaptation and risk management.
213
Frequently Asked Questions
FAQ 12.2: What are climatic thresholds and why are they important?
Many systems have thresholds that can lead to sudden changes, if climate conditions exceed them.
Adaptation and risk management efforts can increase overall system resilience by identifying and
changing tolerance thresholds.
FAQ 12.2, Figure 1 | Crop response to maximum temperature thresholds. Crop growth rate responds to daily maximum temperature increases,
leading to reduced growth and crop failure as temperatures exceed critical and limiting temperature thresholds, respectively. Note that changes in other
environmental factors (such as carbon dioxide and water) may increase the tolerance of plants to increasing temperatures.
FAQ
214
Frequently Asked Questions
FAQ 12.3 | How Will Climate Change Affect the Regional Characteristics of a Climate Hazard?
Human-driven climate change can alter the regional characteristics of a climate hazard by changing the magnitude
or intensity of the climate hazard, the frequency with which it occurs, the duration that hazardous conditions
persist, the timing when the hazard occurs, or the spatial extent threatened by the hazard. By examining each
of these aspects of a hazard’s profile change, climate services may provide climate risk information that allows
decision makers to better tailor adaptation, mitigation and risk management strategies.
A climate hazard is a climate condition with the potential to harm natural systems or society. Examples include
heatwaves, droughts, heavy snowfall events and sea level rise. Climate scientists look for patterns in climatic
impact-drivers to detect the signature of changing hazards that may influence stakeholder planning (FAQ 12.1).
Climate service providers work with stakeholders and impacts experts to identify key system responses and
tolerance thresholds (FAQ 12.2) and then examine historical observations and future climate projections to
identify associated changes to the characteristics of a regional hazard’s profile. Climate change can alter at least
five different characteristics of the hazard profile of a region (FAQ 12.3, Figure 1):
Magnitude or intensity is the raw value of a climate hazard, such as an increase in the maximum yearly temperature
or in the height of flooding that results from a coastal storm with a 1% change of occurring each year.
Frequency is the number of times that a climate hazard reaches or surpasses a threshold over a given period.
For example, increases to the number of heavy snowfall events, tornadoes, or floods experienced in a year or in
a decade.
Duration is the length of time over which hazardous conditions persist beyond a threshold, such as an increase
in the number of consecutive days where maximum air temperature exceeds 35°C, the number of consecutive
months of drought conditions, or the number of days that a tropical cyclone affects a location.
Timing captures the occurrence of a hazardous event in relation to the course of a day, season, year, or other
period in which sectoral elements are evolving or co-dependent (such as the time of year when migrating animals
expect to find a seasonal food supply). Examples include a shift towards an earlier day of the year when the last
spring frost occurs or a delay in the typical arrival date for the first seasonal rains, the length of the winter period FAQ
when the ground is typically covered by snow, or a reduction in the typical time needed for soil moisture to move
from normal to drought conditions.
Spatial extent is the region in which a hazardous condition is expected, such as the area currently threatened by
tropical cyclones, geographical areas where the coldest day of the year restricts a particular pest or pathogen,
terrain where permafrost is present, the area that would flood following a common storm, zones where climate
conditions are conducive to outdoor labour, or the size of a marine heatwave.
Hazard profile changes are often intertwined or stem from related physical changes to the climate system. For
example, changes in the frequency and magnitude of extreme events are often directly related to each other
as a result of atmospheric dynamics and chemical processes. In many cases, one aspect of hazard change is more
apparent than others, which may provide a first emergent signal indicating a larger set of changes to come
(FAQ 1.2).
Information about how a hazard has changed or will change helps stakeholders prioritize more robust adaptation,
mitigation and risk management strategies. For example, allocation of limited disaster relief resources may be
designed to recognize that tropical cyclones are projected to become more intense even as the frequency of
those storms may not change. Planning may also factor in the fact that even heatwaves that are not record-
breaking in their intensity can still be problematic for vulnerable populations when they persist over a long
period. Likewise, firefighters recognize new logistical challenges in the lengthening of the fire weather season
and an expansion of fire conditions into parts of the world where fires were not previously a great concern.
Strong engagement between climate scientists and stakeholders therefore helps climate services tailor and
communicate clear information about the types of changing climate hazards to be addressed in resilience efforts.
215
Frequently Asked Questions
H H
H
H River
Time Time
FAQ
FAQ 12.3, Figure 1 | Types of changes to a region’s hazard profile. The first five panels illustrate how climate changes can alter a hazard’s intensity
(or magnitude), frequency, duration, and timing (by seasonality and speed of onset) in relation to a hazard threshold (horizontal grey line, marked ‘H’). The
difference between the historical climate (blue) and future climate (red) shows the changing aspects of climate change that stakeholders will have to manage.
The bottom right-hand panel shows how a given climate hazard (such as a current once-in-100-year river flood, geographic extent in blue) may reach new
geographical areas under a future climate change (extended area in red).
216
Glossary
217
Glossary
Coordinating Editors:
J.B. Robin Matthews (France/United Kingdom), Vincent Möller (Germany), Renée van Diemen
(The Netherlands/United Kingdom), Jan S. Fuglestvedt (Norway), Valérie Masson-Delmotte
(France), Carlos Méndez (Venezuela), Sergey Semenov (Russian Federation), Andy Reisinger
(New Zealand)
Editorial Team:
Rondrotiana Barimalala (South Africa/Madagascar), Roxana Bojariu (Romania), Annalisa Cherchi
(Italy), Peter M. Cox (United Kingdom), Sergio Henrique Faria (Spain/Brazil), Piers Forster
(United Kingdom), Christopher Jones (United Kingdom), Nana Ama Browne Klutse (Ghana),
Charles Koven (United States of America), Svitlana Krakovska (Ukraine), Sawsan K. Mustafa (Sudan),
Friederike Otto (United Kingdom/Germany), Matthew D. Palmer (United Kingdom), Tamzin Palmer
(United Kingdom), Wilfried Pokam Mba (Cameroon), Roshanka Ranasinghe (The Netherlands/Sri
Lanka, Australia), Pedro Scheel Monteiro (South Africa), Joeri Rogelj (United Kingdom/ Belgium),
Sharon L. Smith (Canada), Ying Sun (China), Andrew Turner (United Kingdom), Bart van den Hurk
(The Netherlands), Émilie Vanvyve (United Kingdom/Belgium), Martin Wild (Switzerland),
Cunde Xiao (China), Prodromos Zanis (Greece)
Note:
This glossary defines some specific terms as the Lead Authors intend them to be interpreted in
the context of this report. Italicized words in definitions indicate that the italicized term is defined
in the Glossary.
219
Annex VII Glossary
1.5°C pathway See Pathways. or ocean, while convection describes the predominantly vertical,
locally induced motions.
Ablation (of glaciers, ice sheets, or snow cover) See Mass
balance/budget (of glaciers or ice sheets). Aerosol A suspension of airborne solid or liquid particles, with
typical particle size in the range of a few nanometres to several
Abrupt change A change in the system that is substantially
tens of micrometres and atmospheric lifetimes of up to several
faster than the typical rate of the changes in its history. See also
days in the troposphere and up to years in the stratosphere. The
Abrupt climate change and Tipping point.
term aerosol, which includes both the particles and the suspending
Abrupt climate change A large-scale abrupt change in the gas, is often used in this report in its plural form to mean ‘aerosol
climate system that takes place over a few decades or less, persists particles’. Aerosols may be of either natural or anthropogenic origin
(or is anticipated to persist) for at least a few decades and causes in the troposphere; stratospheric aerosols mostly stem from volcanic
substantial impacts in human and/or natural systems. See also Abrupt eruptions. Aerosols can cause an effective radiative forcing directly
change and Tipping point. through scattering and absorbing radiation (aerosol–radiation
interaction), and indirectly by acting as cloud condensation nuclei or
Accumulation (of glaciers, ice sheets, or snow cover) See ice nucleating particles that affect the properties of clouds (aerosol–
Mass balance/budget (of glaciers or ice sheets). cloud interaction), and upon deposition on snow- or ice-covered
Active layer Layer of ground above permafrost subject to annual surfaces. Atmospheric aerosols may be either emitted as primary
thawing and freezing. particulate matter or formed within the atmosphere from gaseous
precursors (secondary production). Aerosols may be composed of
Adaptation In human systems, the process of adjustment to sea salt, organic carbon, black carbon (BC), mineral species (mainly
actual or expected climate and its effects, in order to moderate harm desert dust), sulphate, nitrate and ammonium or their mixtures. See
or exploit beneficial opportunities. In natural systems, the process also Short-lived climate forcers (SLCFs).
of adjustment to actual climate and its effects; human intervention
may facilitate adjustment to expected climate and its effects.
See also Adaptation options, Adaptive capacity and Maladaptive Aerosol effective radiative forcing (ERFari+aci) See
actions (Maladaptation). Aerosol–radiation interaction.
Adaptation options The array of strategies and measures Aerosol optical depth (AOD) Wavelength-dependent aerosol
that are available and appropriate for addressing adaptation. They optical depth is a measure of the aerosol contribution to extinction of
include a wide range of actions that can be categorized as structural, top-of-the-atmosphere solar intensity measured at the ground. AOD
institutional, ecological or behavioural. is unitless.
Adaptive capacity The ability of systems, institutions, humans Fine-mode aerosol optical depth
and other organisms to adjust to potential damage, to take advantage Aerosol optical depth due to aerosol particles smaller than 1 μm
of opportunities, or to respond to consequences (MA, 2005). in radius.
Added value Improvement of the representation of some climatic Aerosol–cloud interaction A process by which a perturbation
aspects by one methodology compared to another methodology. For to aerosol affects the microphysical properties and evolution of
instance, downscaling a coarse resolution global climate model may clouds through the aerosol role as cloud condensation nuclei or ice
improve the representation of regional climate in complex terrain. nuclei, particularly in ways that affect radiation or precipitation; such
processes can also include the effect of clouds and precipitation on
Adjustments (in relation to effective radiative aerosol. The aerosol perturbation can be anthropogenic or come from
forcing) The response to an agent perturbing the climate system some natural source. The radiative forcing from such interactions has
that is driven directly by the agent, independently of any change traditionally been attributed to numerous indirect aerosol effects,
in global surface temperature. For example, carbon dioxide and but in this report, only two levels of radiative forcing (or effect)
aerosols, by altering internal heating and cooling rates within the are distinguished:
atmosphere, can each cause changes to cloud cover and other
variables thereby producing an effective radiative forcing even in Effective radiative forcing (or effect) due to aerosol–cloud interactions
the absence of any surface warming or cooling. Adjustments are (ERFaci)
usually rapid in the sense that they begin to occur right away, The final radiative forcing (or effect) from the aerosol perturbation,
before climate feedbacks which are driven by global surface including the adjustments to the initial change in droplet or crystal
warming (although some adjustments may still take significant time formation rate. These adjustments include changes in the strength of
to proceed to completion, for example those involving vegetation convection, precipitation efficiency, cloud fraction, lifetime or water
or ice sheets). content of clouds, and the formation or suppression of clouds in
remote areas due to altered circulations.
Adjustment time See Response time or adjustment time.
Instantaneous radiative forcing (or effect) due to aerosol–cloud
Advection Transport of water or air along with its properties interactions (IRFaci)
(e.g., temperature, chemical tracers) by winds or currents. Regarding The radiative forcing (or radiative effect, if the perturbation is
AVII the general distinction between advection and convection, the internally generated) due to the change in number or size distribution
former describes transport by large-scale motions of the atmosphere of cloud droplets or ice crystals that is the proximate result of an
220
Glossary Annex VII
aerosol perturbation, with other variables (in particular total cloud Air pollution Degradation of air quality with negative effects
water content) remaining equal. In liquid clouds, an increase in cloud on human health or the natural or built environment due to the
droplet concentration and surface area would increase the cloud introduction, by natural processes or human activity, into the
albedo. This effect is also known as the cloud albedo effect, first atmosphere of substances (gases, aerosol) which have a direct
indirect effect, or Twomey effect. It is a largely theoretical concept (primary pollutants) or indirect (secondary pollutants) harmful effect.
that cannot readily be isolated in observations or comprehensive See also Short-lived climate forcers (SLCFs).
process models due to the ubiquity of adjustments.
Airborne fraction The fraction of total carbon dioxide (CO2)
See also Aerosol–radiation interaction. emissions (from fossil fuels and land-use change) remaining in
the atmosphere.
Aerosol–radiation interaction An interaction of aerosol
directly with radiation produces radiative effects. In this report, two Albedo The proportion of sunlight (solar radiation) reflected by
levels of radiative forcing (or effect) are distinguished: a surface or object, often expressed as a percentage. Clouds, snow
and ice usually have high albedo; soil surfaces cover the albedo range
Aerosol effective radiative forcing (ERFari+aci)
from high to low; vegetation in the dry season and/or in arid zones
The total effective radiative forcing due to both aerosol–cloud and
can have high albedo, whereas photosynthetically active vegetation
aerosol–radiation interactions is denoted aerosol effective radiative
and the ocean have low albedo. The Earth’s planetary albedo changes
forcing (ERFari+aci).
mainly through changes in cloudiness and of snow, ice, leaf area and
Effective radiative forcing (or effect) due to aerosol–radiation land cover.
interactions (ERFari)
Alkalinity See Total alkalinity.
The final radiative forcing (or effect) from the aerosol perturbation,
including adjustments to the initial change in radiation. These Altimetry A technique for measuring the height of the Earth’s
adjustments include changes in cloud caused by the impact of surface with respect to the geocentre of the Earth within a defined
the radiative heating on convective or larger-scale atmospheric terrestrial reference frame (geocentric sea level). See also Geocentric
circulations, traditionally known as semi-direct aerosol forcing sea level change.
(or effect).
Annular modes Hemispheric scale patterns of atmospheric
Instantaneous radiative forcing (or effect) due to aerosol–radiation variability characterized by opposing and synchronous fluctuations
interactions (IRFari) in sea level pressure between the polar caps and mid-latitudes,
The radiative forcing (or radiative effect, if the perturbation is with a structure exhibiting a high degree of zonal symmetry, and
internally generated) of an aerosol perturbation due directly to with no real preferred time scales ranging from days to decades. In
aerosol–radiation interactions, with all environmental variables each hemisphere, these fluctuations reflect changes in the latitudinal
remaining unaffected. Traditionally known in the literature as the position and strength of the mid-latitude jets and associated storm
direct aerosol forcing (or effect). tracks. Annular modes are defined as the leading mode of variability
of extratropical sea level pressure or geopotential heights and are
See also Aerosol–cloud interaction.
known as the Northern Annular Mode (NAM) and Southern Annular
Afforestation Conversion to forest of land that historically has Mode (SAM) in the two hemispheres, respectively.
not contained forests. [Note: For a discussion of the term forest and
Northern Annular Mode (NAM)
related terms such as afforestation, reforestation and deforestation,
A see-saw latitudinal fluctuation in Northern Hemisphere sea level
see the 2006 IPCC Guidelines for National Greenhouse Gas Inventories
pressure or geopotential height between the Arctic and the mid-
and their 2019 Refinement, and information provided by the United
latitudes. The NAM has some links with the stratospheric polar
Nations Framework Convention on Climate Change (IPCC, 2006,
vortex and is related to the fluctuation in strength and latitude of the
2019; UNFCCC, 2021a, b).] See also Deforestation, Reforestation,
mean westerlies. Its variance is maximum in winter and its pattern
Anthropogenic removals and Carbon dioxide removal (CDR).
has a strong regional expression in the North Atlantic, being strongly
Agreement In this Report, the degree of agreement within the correlated with the North Atlantic Oscillation index. The NAM is also
scientific body of knowledge on a particular finding is assessed known as the Arctic Oscillation (AO). In its positive phase, the NAM is
based on multiple lines of evidence (e.g., mechanistic understanding, characterized by anomalously low pressure over the Arctic and high
theory, data, models, expert judgement) and expressed qualitatively pressure over the mid-latitudes/subtropics, with a strengthening of
(Mastrandrea et al., 2010). See also Confidence, Likelihood, the zonally averaged westerly winds on their polar flank that confines
Uncertainty and Evidence. colder air across the Arctic. The negative NAM phase is characterized
by a more distorted wind pattern and jet meanders that increase
Agricultural and ecological drought See Drought.
storminess in the mid-latitude regions. See Section AIV.2.1 in Annex
Air mass A widespread body of air, the approximately IV of the AR6 WGI report.
homogeneous properties of which (i) have been established while
Southern Annular Mode (SAM)
that air was situated over a particular region of the Earth’s surface,
The leading mode of climate variability of Southern Hemisphere sea
and (ii) undergo specific modifications while in transit away from the
level pressure and geopotential height, which is associated with the
source region (AMS, 2021).
strength and latitudinal shifts in the mid- to high-latitudes westerly AVII
wind belt. The SAM is also known as the Antarctic Oscillation (AAO).
221
Annex VII Glossary
A positive SAM phase is defined as lower-than-normal pressures in this report. As a consequence, the land-related net GHG emission
over the polar regions and higher-than-normal pressures in the estimates from global models included in this report are not necessarily
southern mid-latitudes, with a contraction towards Antarctica and directly comparable with land use, land-use change and forestry
strengthening of the westerly wind belt. The negative SAM phase (LULUCF) estimates in national GHG Inventories.] See also Carbon
exhibits positive high-latitude pressure anomalies, negative mid- dioxide removal (CDR), Afforestation, Enhanced weathering, Ocean
latitude pressure anomalies and a weaker westerly flow expanded alkalinization/Ocean alkalinity enhancement and Reforestation.
towards the equator. See Section AIV.2.2 in Annex IV of the AR6 WGI
Anthropogenic subsidence Downward motion of the land
report. See also Annular modes.
surface induced by anthropogenic drivers (e.g., loading, extraction
Anomaly The deviation of a variable from its value averaged over of hydrocarbons and/or groundwater, drainage, mining activities)
a reference period. causing sediment compaction or subsidence/deformation of the
sedimentary sequence, or oxidation of organic material, thereby
Antarctic amplification See Polar amplification.
leading to relative sea level rise.
Antarctic Ice Sheet (AIS) See Ice sheet.
Apparent hydrological sensitivity (ηa) The change in global
Antarctic oscillation (AAO) See Southern Annular Mode (SAM) mean precipitation per degree Celsius of global mean surface
(under Annular modes). air temperature (GSAT) change with units of % per °C, although
it can also be calculated as W m–2 per °C. See also Hydrological
Anthropocene A proposed new geological epoch resulting from
sensitivity (η).
significant human-driven changes to the structure and functioning of
the Earth System, including the climate system. Originally proposed Arctic amplification See Polar amplification.
in the Earth system science community in 2000, the proposed new
Arctic oscillation (AO) See Northern Annular Mode (NAM)
epoch is undergoing a formalization process within the geological
(under Annular modes).
community based on the stratigraphic evidence that human activities
have changed the Earth system to the extent of forming geological Arid zone Areas where vegetation growth is severely constrained
deposits with a signature that is distinct from those of the Holocene, due to limited water availability. For the most part, the native
and which will remain in the geological record. Both the stratigraphic vegetation of arid zones is sparse. There is high rainfall variability,
and Earth system approaches to defining the Anthropocene consider with annual averages below 300 mm. Crop farming in arid zones
the mid-20th century to be the most appropriate starting date requires irrigation.
(Steffen et al., 2016), although others have been proposed and
Aridity The state of a long-term climatic feature characterized
continue to be discussed. The Anthropocene concept has already
by low average precipitation or available water in a region. Aridity
been informally adopted by diverse disciplines and the public to
generally arises from widespread persistent atmospheric subsidence
denote the substantive influence of humans on the Earth system.
or anticyclonic conditions, and from more localized subsidence in the
Anthropogenic Resulting from or produced by human activities. lee side of mountains (adapted from Gbeckor-Kove, 1989; Türkeş,
1999). See also Drought.
Anthropogenic emissions Emissions of greenhouse gases
(GHGs), precursors of GHGs and aerosols caused by human activities. Artificial ocean upwelling (AOUpw) A potential carbon
These activities include the burning of fossil fuels, deforestation, land dioxide removal (CDR) method that aims to artificially pump up
use and land-use changes (LULUC), livestock production, fertilization, cooler, nutrient-rich waters from deep in the ocean to the surface.
waste management, and industrial processes. See also Anthropogenic The aim is to stimulate phytoplankton activity and thereby increase
and Anthropogenic removals. ocean CO2 uptake.
Anthropogenic removals The withdrawal of greenhouse gases Assets Natural or human-made resources that provide current
(GHGs) from the atmosphere as a result of deliberate human or future utility, benefit, economic or intrinsic value to natural or
activities. These include enhancing biological sinks of CO2 and using human systems.
chemical engineering to achieve long-term removal and storage.
Atlantic Equatorial Mode See Atlantic Zonal Mode (AZM)
Carbon dioxide capture and storage (CCS), which alone does not
under Tropical Atlantic Variability (TAV).
remove CO2 from the atmosphere, can help reduce atmospheric CO2
from industrial and energy-related sources if it is combined with Atlantic Meridional Mode (AMM) See Tropical Atlantic
bioenergy production (BECCS), or if CO2 is captured from the air Variability (TAV).
directly and stored (DACCS). [Note: In the 2006 IPCC Guidelines for
Atlantic Meridional Overturning Circulation (AMOC) See
national GHG Inventories (IPCC, 2006), which are used in reporting of
Meridional overturning circulation (MOC).
emissions to the UNFCCC, ‘anthropogenic’ land-related GHG fluxes
are defined as all those occurring on ‘managed land’, i.e., ‘where Atlantic Multi-decadal Oscillation (AMO) See Atlantic Multi-
human interventions and practices have been applied to perform decadal Variability (AMV).
production, ecological or social functions’. However, some removals
(e.g., removals associated with CO2 fertilization and N deposition) Atlantic Multi-decadal Variability (AMV) Large-scale
are not considered as ‘anthropogenic’, or are referred to as ‘indirect’ fluctuations observed from one decade to the next in a variety of
AVII instrumental records and proxy reconstructions over the entire
anthropogenic effects, in some of the scientific literature assessed
North Atlantic ocean and surrounding continents. Fingerprints of
222
Glossary Annex VII
AMV can be found at the surface ocean, which is characterized by Autotrophic respiration Respiration by photosynthetic (see
swings in basin-scale sea surface temperature anomalies reflecting photosynthesis) organisms (e.g., plants and algae).
the interaction with the atmosphere. The positive phase of the
Avalanche A mass of snow, ice, earth or rocks, or a mixture of
AMV is characterized by anomalous warming over the entire North
these, falling down a mountainside.
Atlantic, with the strongest amplitude in the subpolar gyre and along
sea ice margin zones in the Labrador Sea and Greenland/Barents Barystatic See Sea level change (sea level rise/sea level fall).
Sea and in the subtropical North Atlantic basin to a lower extent.
Basal lubrication Reduction of friction at the base of an ice sheet
In the AR6 WGI report, the term AMV is preferred to Atlantic Multi-
or glacier due to lubrication by meltwater. This can allow the glacier
decadal Oscillation (AMO) used in previous IPCC reports because
or ice sheet to slide over its base. Meltwater may be produced by
there is no preferred time scale of decadal variability as the term
pressure-induced melting, friction or geothermal heat, or surface
oscillation would indirectly imply. See Section AIV.2.7 in Annex IV of
melt may drain to the base through holes in the ice.
the AR6 WGI report.
Baseline/reference See Reference scenario (under Scenario) and
Atlantic Niño See Atlantic Zonal Mode (AZM) under Tropical
Reference period.
Atlantic Variability (TAV).
Baseline scenario See Reference scenario (under Scenario).
Atlantic Zonal Mode (AZM) See Tropical Atlantic Variability
(TAV). Bifurcation point See Tipping point.
Atmosphere The gaseous envelope surrounding the Earth, Biodiversity Biodiversity or biological diversity means the
divided into five layers – the troposphere which contains half of variability among living organisms from all sources including, among
the Earth’s atmosphere, the stratosphere, the mesosphere, the other things, terrestrial, marine and other aquatic ecosystems, and the
thermosphere and the exosphere, which is the outer limit of the ecological complexes of which they are part; this includes diversity
atmosphere. The dry atmosphere consists almost entirely of nitrogen within species, between species, and of ecosystems (UN, 1992). See
(78.1% volume mixing ratio) and oxygen (20.9% volume mixing also Ecosystem.
ratio), together with a number of trace gases, such as argon (0.93%
volume mixing ratio), helium and radiatively active greenhouse gases Bioenergy with carbon dioxide capture and storage
(GHGs) such as carbon dioxide (CO2) (0.04% volume mixing ratio), (BECCS) Carbon dioxide capture and storage (CCS) technology
methane (CH4 ), nitrous oxide (N2O) and ozone (O3). In addition, applied to a bioenergy facility. Note that depending on the total
the atmosphere contains the GHG water vapour (H2O), whose emissions of the BECCS supply chain, carbon dioxide (CO2) can be
concentrations are highly variable (0–5% volume mixing ratio) as the removed from the atmosphere. See also Carbon dioxide capture
sources (evapotranspiration) and sinks (precipitation) of water vapour and storage (CCS), Anthropogenic removals and Carbon dioxide
show large spatio-temporal variations, and atmospheric temperature removal (CDR).
exerts a strong constraint on the amount of water vapour an air Biogenic volatile organic compounds (BVOCs) See Volatile
parcel can hold. The atmosphere also contains clouds and aerosols. organic compounds (VOCs).
See also Hydrological cycle, Stratosphere and Troposphere.
Biogeophysical potential See Mitigation potential.
Atmosphere–ocean general circulation model
(AOGCM) See General circulation model (GCM). Biological (carbon) pump A series of ocean processes through
which inorganic carbon (as carbon dioxide, CO2) is fixed as organic
Atmospheric boundary layer The atmospheric layer adjacent matter by photosynthesis in sunlit surface water and then transported
to the Earth’s surface that is affected by friction against that boundary to the ocean interior, and possibly the sediment, resulting in the
surface, and possibly by transport of heat and other variables across storage of carbon.
that surface (AMS, 2021). The lowest 100 m of the boundary layer
(about 10% of the boundary layer thickness), where mechanical Biomass Organic material excluding the material that is fossilized
generation of turbulence is dominant, is called the surface boundary or embedded in geological formations. Biomass may refer to the
layer or surface layer. mass of organic matter in a specific area (ISO, 2014).
Atmospheric lifetime See Lifetime. Biosphere (terrestrial and marine) The part of the Earth
system comprising all ecosystems and living organisms, in the
Atmospheric rivers (ARs) Long, narrow (up to a few hundred atmosphere, on land (terrestrial biosphere) or in the oceans (marine
km wide), shallow (up to a few km deep) and transient corridors of biosphere), including derived dead organic matter, such as litter, soil
strong horizontal water vapour transport that are typically associated organic matter and oceanic detritus.
with a low-level jet stream ahead of the cold front of an extratropical
cyclone (ETC) (Ralph et al., 2018). Bipolar seesaw (also inter-hemispheric seesaw, inter-
hemispheric asymmetry, hemispheric asymmetry) A
Attribution Attribution is defined as the process of evaluating phenomenon in which temperature changes in the Northern and
the relative contributions of multiple causal factors to a change or Southern hemispheres are related but out of phase, generally
event with an assessment of confidence. inferred to represent a change in the magnitude or sign of net
Australian and Maritime Continent monsoon heat transport across the equator. Originally called hemispheric AVII
(AusMCM) See Global monsoon. asymmetry and linked to changes in thermohaline overturning
223
Annex VII Glossary
circulation on multi-millennial scales (Mix et al., 1986), later named Calcification The process of biologically precipitating calcium
bipolar seesaw and applied to millennial scales (Broecker, 1998) with carbonate minerals to create organism shells, skeletons, otoliths, or
a similar thermohaline mechanism (Stocker and Johnsen, 2003). other body structures. The chemical equation describing calcification
See also Meridional overturning circulation (MOC) and Deglacial or is Ca2+ (aq) + 2HCO3– (aq) → CaCO3(s) + CO2 + H2O. Aragonite and
deglaciation or glacial termination. calcite are two common crystalline forms of biologically precipitated
calcium carbonate minerals that have different solubilities.
Black carbon (BC) A relatively pure form of carbon, also known
as soot, arising from the incomplete combustion of fossil fuels, biofuel, Calving (of glaciers or ice sheets) The breaking off of
and biomass. It only stays in the atmosphere for days or weeks. BC discrete pieces of ice from a glacier, ice sheet or an ice shelf into
is a climate forcing agent with strong warming effect, both in the lake or seawater, producing icebergs. This is a form of mass loss from
atmosphere and when deposited on snow or ice. See also Aerosol an ice body.
and Atmosphere.
Canopy temperature The temperature within the canopy of
Blocking Associated with persistent, slow-moving high-pressure a vegetation structure.
systems that obstruct the prevailing westerly winds in the middle
Carbon budget Refers to two concepts in the literature: (i)
and high latitudes and the normal eastward progress of extratropical
an assessment of carbon cycle sources and sinks on a global
transient storm systems. It is an important component of the intra-
level, through the synthesis of evidence for fossil fuel and cement
seasonal climate variability in the extratropics and can cause
emissions, emissions and removals associated with land use and
long-lived weather conditions such as cold spells in winter and
land-use change, ocean and natural land sources and sinks of
summer heatwaves.
carbon dioxide (CO2), and the resulting change in atmospheric CO2
Blue carbon Biologically driven carbon fluxes and storage in concentration. This is referred to as the global carbon budget; (ii)
marine systems that are amenable to management. Coastal blue the maximum amount of cumulative net global anthropogenic CO2
carbon focuses on rooted vegetation in the coastal zone, such as tidal emissions that would result in limiting global warming to a given
marshes, mangroves and seagrasses. These ecosystems have high level with a given probability, taking into account the effect of other
carbon burial rates on a per unit area basis and accumulate carbon anthropogenic climate forcers. This is referred to as the total carbon
in their soils and sediments. They provide many non-climatic benefits budget when expressed starting from the pre-industrial period, and
and can contribute to ecosystem-based adaptation. If degraded or as the remaining carbon budget when expressed from a recent
lost, coastal blue carbon ecosystems are likely to release most of their specified date.
carbon back to the atmosphere. There is current debate regarding
Note 1: Net anthropogenic CO2 emissions are anthropogenic CO2
the application of the blue carbon concept to other coastal and non-
emissions minus anthropogenic CO2 removals. See also Carbon
coastal processes and ecosystems, including the open ocean. See
dioxide removal (CDR).
also Sequestration.
Note 2: The maximum amount of cumulative net global anthropogenic
Brewer–Dobson circulation The meridional overturning
CO2 emissions is reached at the time that annual net anthropogenic
circulation of the stratosphere transporting air upward in the tropics,
CO2 emissions reach zero.
poleward to the winter hemisphere, and downward at polar and
subpolar latitudes. The Brewer–Dobson circulation is driven by the Note 3: The degree to which anthropogenic climate forcers other than
interaction between upward propagating planetary waves and the CO2 affect the total carbon budget and remaining carbon budget
mean flow. depends on human choices about the extent to which these forcers
are mitigated and their resulting climate effects.
Burden The total mass of a substance of concern in the atmosphere.
Note 4: The notions of a total carbon budget and remaining carbon
Business as usual (BAU) The term business as usual scenario
budget are also being applied in parts of the scientific literature and
has been used to describe a scenario that assumes no additional
by some entities at regional, national, or sub-national levels. The
policies beyond those currently in place and that patterns of socio-
distribution of global budgets across individual different entities
economic development are consistent with recent trends. The term
and emitters depends strongly on considerations of equity and other
is now used less frequently than in the past. See also Reference
value judgements.
scenario (under Scenario).
Carbon cycle The flow of carbon (in various forms, e.g., as
13
C Stable isotope of carbon having an atomic weight of
carbon dioxide (CO2), carbon in biomass, and carbon dissolved in
approximately 13. Measurements of the ratio of 13C/12C in carbon
the ocean as carbonate and bicarbonate) through the atmosphere,
dioxide (CO2) molecules are used to infer the importance of different
hydrosphere, terrestrial and marine biosphere and lithosphere. In this
carbon cycle and climate processes and the size of the terrestrial
report, the reference unit for the global carbon cycle is GtCO2 or GtC
carbon reservoir.
(one Gigatonne = 1 Gt = 1015 grams; 1 GtC corresponds to 3.664
14
C Unstable isotope of carbon having an atomic weight of GtCO2). See also Ocean carbon cycle.
approximately 14 and a half-life of about 5700 years. It is often used
Carbon dioxide (CO2) A naturally occurring gas, CO2 is also
for dating purposes going back some 40 kyr. Its variation in time is
a by-product of burning fossil fuels (such as oil, gas and coal), of
affected by the magnetic fields of the Sun and Earth, which influence
AVII burning biomass, of land-use change (LUC) and of industrial
its production from cosmic rays.
processes (e.g., cement production). It is the principal anthropogenic
224
Glossary Annex VII
greenhouse gas (GHG) that affects the Earth’s radiative balance. It Carbon source See Source.
is the reference gas against which other GHGs are measured and
Carbon–climate feedback See Climate–carbon cycle feedback.
therefore has a global warming potential (GWP) of 1.
Carbonaceous aerosol Aerosol consisting predominantly of
Carbon dioxide (CO2) fertilization The increase of plant
organic substances and black carbon.
photosynthesis and water-use efficiency in response to increased
atmospheric carbon dioxide (CO2) concentration. Whether this Carbonate counter pump See Carbonate pump.
increased photosynthesis translates into increased plant growth
Carbonate pump Ocean carbon fixation through the biological
and carbon storage on land depends on the interacting effects of
formation of carbonates, primarily by plankton that generate
temperature, moisture and nutrient availability.
bio-mineral particles that sink to the ocean interior, and possibly
Carbon dioxide capture and storage (CCS) A process in the sediment. It is also called carbonate counter-pump, since the
which a relatively pure stream of carbon dioxide (CO2) from industrial formation of calcium carbonate (CaCO3) is accompanied by the release
and energy-related sources is separated (captured), conditioned, of carbon dioxide (CO2) to surrounding water and subsequently to the
compressed and transported to a storage location for long-term atmosphere.
isolation from the atmosphere. Sometimes referred to as carbon
Catchment An area that collects and drains precipitation.
capture and storage. See also Bioenergy with carbon dioxide capture
and storage (BECCS), Sequestration, Anthropogenic removals and Cenozoic Era The third and current geological Era, which began
Carbon dioxide removal (CDR). 66.0 Ma. It comprises the Paleogene, Neogene and Quaternary
Periods.
Carbon dioxide removal (CDR) Anthropogenic activities
removing carbon dioxide (CO2) from the atmosphere and durably Central Pacific El Niño See El Niño–Southern Oscillation (ENSO).
storing it in geological, terrestrial, or ocean reservoirs, or in products.
It includes existing and potential anthropogenic enhancement of Chaotic A dynamical system such as the climate system, governed
biological or geochemical CO2 sinks and direct air carbon dioxide by non-linear deterministic equations, may exhibit erratic or chaotic
capture and storage (DACCS), but excludes natural CO2 uptake behaviour in the sense that very small changes in the initial state of
not directly caused by human activities. See also Anthropogenic the system lead to large and apparently unpredictable changes in its
removals, Afforestation, Enhanced weathering, Ocean alkalinization/ temporal evolution. Such chaotic behaviour limits the predictability
Ocean alkalinity enhancement, Reforestation, Bioenergy with carbon of the state of a non-linear dynamical system at specific future times,
dioxide capture and storage (BECCS) and Carbon dioxide capture and although changes in its statistics may still be predictable given
storage (CCS). changes in the system parameters or boundary conditions.
Carbon neutrality Condition in which anthropogenic CO2 Charcoal Material resulting from charring of biomass, usually
emissions associated with a subject are balanced by anthropogenic retaining some of the microscopic texture typical of plant tissues;
CO2 removals. The subject can be an entity such as a country, chemically it consists mainly of carbon with a disturbed graphitic
an organization, a district or a commodity, or an activity such as structure, with lesser amounts of oxygen and hydrogen.
a service and an event. Carbon neutrality is often assessed over the life Chlorofluorocarbons (CFCs) An organic compound that
cycle including indirect (‘scope 3’) emissions, but can also be limited contains chlorine, carbon, hydrogen and fluorine and is used for
to the emissions and removals, over a specified period, for which the refrigeration, air conditioning, packaging, plastic foam, insulation,
subject has direct control, as determined by the relevant scheme. solvents or aerosol propellants. Because they are not destroyed in
Note 1: Carbon neutrality and net zero CO2 emissions are overlapping the lower atmosphere, CFCs drift into the upper atmosphere where,
concepts. The concepts can be applied at global or sub-global scales given suitable conditions, they lead to ozone (O3 ) depletion. They
(e.g., regional, national and sub-national). At a global scale, the terms are some of the greenhouse gases (GHGs) covered under the 1987
carbon neutrality and net zero CO2 emissions are equivalent. At sub- Montreal Protocol, as a result of which manufacturing of these
global scales, net zero CO2 emissions is generally applied to emissions gases has been phased out, and they are being replaced by other
and removals under direct control or territorial responsibility of the compounds, including hydrofluorocarbons (HFCs).
reporting entity, while carbon neutrality generally includes emissions Chronology Arrangement of events according to dates or times
and removals within and beyond the direct control or territorial of occurrence.
responsibility of the reporting entity. Accounting rules specified by
GHG programmes or schemes can have a significant influence on the Cirrus cloud thinning (CCT) See Solar radiation modification
quantification of relevant CO2 emissions and removals. (SRM).
Note 2: In some cases, achieving carbon neutrality may rely on the Clathrate (methane) A partly frozen slushy mix of methane gas
supplementary use of offsets to balance emissions that remain after and ice, usually found in sediments.
actions by the reporting entity are taken into account. Clausius–Clapeyron equation/relationship The thermodynamic
See also Greenhouse gas neutrality and Net zero CO2 emissions. relationship between temperature and the vapour pressure of
a substance in which two phases of the substance are in equilibrium
Carbon sequestration See Sequestration. (e.g., liquid water and water vapour). For gases such as water vapour, AVII
Carbon sink See Sink.
225
Annex VII Glossary
this relation gives the increase in equilibrium (or saturation) vapour A specific subcategory of zero emissions commitment is the Zero
pressure per unit change in air temperature. CO2 Emissions Commitment which refers to the climate system
response to CO2 emissions after setting these to net zero. The
Climate Climate in a narrow sense is usually defined as the average
CO2-only definition is of specific use in estimating remaining
weather, or more rigorously as the statistical description in terms of
carbon budgets.
the mean and variability of relevant quantities over a period of time
ranging from months to thousands or millions of years. The classical Climate extreme (extreme weather or climate event) The
period for averaging these variables is 30 years, as defined by the occurrence of a value of a weather or climate variable above (or
World Meteorological Organization (WMO). The relevant quantities below) a threshold value near the upper (or lower) ends of the range
are most often surface variables such as temperature, precipitation of observed values of the variable. By definition, the characteristics
and wind. Climate in a wider sense is the state, including a statistical of what is called extreme weather may vary from place to place in an
description, of the climate system. absolute sense. When a pattern of extreme weather persists for some
time, such as a season, it may be classified as an extreme climate
Climate change A change in the state of the climate that can be
event, especially if it yields an average or total that is itself extreme
identified (e.g., by using statistical tests) by changes in the mean and/
(e.g., high temperature, drought, or heavy rainfall over a season). For
or the variability of its properties and that persists for an extended
simplicity, both extreme weather events and extreme climate events
period, typically decades or longer. Climate change may be due to
are referred to collectively as ‘climate extremes’.
natural internal processes or external forcings such as modulations
of the solar cycles, volcanic eruptions and persistent anthropogenic Climate feedback An interaction in which a perturbation in one
changes in the composition of the atmosphere or in land use. Note climate quantity causes a change in a second, and the change in
that the United Nations Framework Convention on Climate Change the second quantity ultimately leads to an additional change in the
(UNFCCC), in its Article 1, defines climate change as: ‘a change of first. A negative feedback is one in which the initial perturbation is
climate which is attributed directly or indirectly to human activity weakened by the changes it causes; a positive feedback is one in
that alters the composition of the global atmosphere and which is which the initial perturbation is enhanced. The initial perturbation
in addition to natural climate variability observed over comparable can either be externally forced or arise as part of internal variability.
time periods’. The UNFCCC thus makes a distinction between climate See also Climate–carbon cycle feedback, Cloud feedback and Ice–
change attributable to human activities altering the atmospheric albedo feedback.
composition and climate variability attributable to natural causes. See
Climate feedback parameter A way to quantify the radiative
also Climate variability, Detection and attribution, Global warming
response of the climate system to a change induced by a radiative
and Ocean acidification (OA).
forcing. It is quantified as the change in net energy flux at the top of
Climate change commitment Climate change commitment atmosphere for a given change in annual global surface temperature.
is defined as the unavoidable future climate change resulting from It has units of W m–2 °C–1.
inertia in the geophysical and socio-economic systems. Different
Climate forecast See Climate prediction.
types of climate change commitment are discussed in the literature
(see subterms). Climate change commitment is usually quantified Climate index A time series constructed from climate variables
in terms of the further change in temperature, but it includes other that provides an aggregate summary of the state of the climate
future changes, for example in the hydrological cycle, in extreme system. For example, the difference between sea level pressure in
weather events, in extreme climate events, and in sea level. Iceland and the Azores provides a simple yet useful historical North
Atlantic Oscillation (NAO) index. Because of their optimal properties,
Constant composition commitment
climate indices are often defined using principal components –
The constant composition commitment is the remaining climate
linear combinations of climate variables at different locations that
change that would result if atmospheric composition, and hence
have maximum variance subject to certain normalization constraints
radiative forcing, were held fixed at a given value. It results from the
(e.g., the Northern Annular Mode (NAM) and Southern Annular
thermal inertia of the ocean and slow processes in the cryosphere
Mode (SAM) indices, which are principal components of Northern
and land surface.
Hemisphere and Southern Hemisphere gridded pressure anomalies,
Constant emissions commitment respectively). Definitions of observational indices for Modes of
The constant emissions commitment is the committed climate change climate variability can be found in Annex IV of the AR6 WGI report.
that would result from keeping anthropogenic emissions constant.
Climate indicator Measures of the climate system, including
Zero emissions commitment large-scale variables and climate proxies. See also Climate metrics.
The zero emissions commitment is an estimate of the subsequent
Key climate indicators
global warming that would result after anthropogenic emissions
Key indicators constitute a finite set of distinct variables that may
are set to zero. It is determined by both inertia in physical climate
collectively point to important overall changes in the climate
system components (ocean, cryosphere, land surface) and carbon
system of broad societal relevance across the atmospheric, oceanic,
cycle inertia. In its widest sense it refers to emissions of each
cryospheric and biospheric domains, with land as an implicit cross-
climate forcer, including greenhouses gases, aerosols and their
cutting theme. Taken together, these indicators would be expected
precursors. The climate response to this can be complex due
AVII to both have changed and continue to change in the future in a
to the different time scale of response of each climate forcer.
226
Glossary Annex VII
coherent and consistent manner. See Cross-Chapter Box 2.2, Table 1 Climate sensitivity The change in the surface temperature
in the AR6 WGI report. in response to a change in the atmospheric carbon dioxide (CO2)
concentration or other radiative forcing. See also Climate feedback
Climate information Information about the past, current
parameter.
state or future of the climate system that is relevant for mitigation,
adaptation and risk management. It may be tailored or ‘co‑produced’ Earth system sensitivity
for specific contexts, taking into account users’ needs and values. The equilibrium surface temperature response of the coupled
atmosphere–ocean–cryosphere–vegetation–carbon cycle system to
Climate metrics Measures of aspects of the overall climate
a doubling of the atmospheric carbon dioxide (CO2) concentration
system response to radiative forcing, such as equilibrium climate
is referred to as Earth system sensitivity. Because it allows ice sheets
sensitivity (ECS), transient climate response (TCR), transient climate
to adjust to the external perturbation, it may differ substantially from
response to cumulative CO2 emissions (TCRE) and the airborne
the equilibrium climate sensitivity derived from coupled atmosphere–
fraction of anthropogenic carbon dioxide. See also Greenhouse gas
ocean models.
emission metric, Climate indicator and Key climate indicators (under
Climate indicator). Effective equilibrium climate sensitivity
An estimate of the surface temperature response to a doubling of
Climate model A qualitative or quantitative representation of
the atmospheric carbon dioxide (CO2) concentration that is evaluated
the climate system based on the physical, chemical and biological
from model output or observations for evolving non-equilibrium
properties of its components, their interactions and feedback
conditions. It is a measure of the strengths of the climate feedbacks at
processes and accounting for some of its known properties. The
a particular time and may vary with forcing history and climate state,
climate system can be represented by models of varying complexity;
and therefore may differ from equilibrium climate sensitivity.
that is, for any one component or combination of components,
a spectrum or hierarchy of models can be identified, differing in such Equilibrium climate sensitivity (ECS)
aspects as the number of spatial dimensions, the extent to which The equilibrium (steady state) change in the surface temperature
physical, chemical or biological processes are explicitly represented, following a doubling of the atmospheric carbon dioxide (CO2)
or the level at which empirical parametrizations are involved. There is concentration from pre-industrial conditions.
an evolution towards more complex models with interactive chemistry
Transient climate response (TCR)
and biology. Climate models are applied as a research tool to study
The surface temperature response for the hypothetical scenario
and simulate the climate and for operational purposes, including
in which atmospheric carbon dioxide (CO2) increases at 1% yr–1
monthly, seasonal and interannual climate predictions. See also Earth
from pre-industrial to the time of a doubling of atmospheric CO2
system model (ESM), Earth system model of intermediate complexity
concentration (year 70).
(EMIC), Energy balance model (EBM), Simple climate model (SCM),
Regional climate model (RCM), Dynamic global vegetation model Transient climate response to cumulative CO2 emissions (TCRE)
(DGVM), General circulation model (GCM) and Emulators. The transient surface temperature change per unit cumulative carbon
dioxide (CO2) emissions, usually 1000 GtC. TCRE combines both
Climate pattern A set of spatially varying coefficients obtained
information on the airborne fraction of cumulative CO2 emissions
by ‘projection’ (regression) of climate variables onto a climate index
(the fraction of the total CO2 emitted that remains in the atmosphere,
time series. When the climate index is a principal component, the
which is determined by carbon cycle processes) and on the transient
climate pattern is an eigenvector of the covariance matrix, referred to
climate response (TCR).
as an empirical orthogonal function (EOF) in climate science.
Climate services Climate services involve the provision of climate
Climate prediction A climate prediction or climate forecast is the
information in such a way as to assist decision-making. The service
result of an attempt to produce (starting from a particular state of the
includes appropriate engagement from users and providers, is based
climate system) an estimate of the actual evolution of the climate in
on scientifically credible information and expertise, has an effective
the future, for example, at seasonal, interannual or decadal time scales.
access mechanism and responds to user needs (Hewitt et al., 2012).
Because the future evolution of the climate system may be highly
sensitive to initial conditions, has chaotic elements and is subject to Climate simulation ensemble A group of parallel model
natural variability, such predictions are usually probabilistic in nature. simulations characterizing historical climate conditions, climate
predictions, or climate projections. Variation of the results across
Climate projection Simulated response of the climate system to
the ensemble members may give an estimate of modelling-based
a scenario of future emissions or concentrations of greenhouse gases
uncertainty. Ensembles made with the same model but different
(GHGs) and aerosols and changes in land use, generally derived using
initial conditions characterize the uncertainty associated with
climate models. Climate projections are distinguished from climate
internal climate variability, whereas multi-model ensembles
predictions by their dependence on the emission/concentration/
including simulations by several models also include the effect of
radiative forcing scenario used, which is in turn based on assumptions
model differences. Perturbed parameter ensembles, in which model
concerning, for example, future socio-economic and technological
parameters are varied in a systematic manner, aim to assess the
developments that may or may not be realized.
uncertainty resulting from internal model specifications within
Climate response A general term for how the climate system a single model. Remaining sources of uncertainty unaddressed with
responds to a radiative forcing. model ensembles are related to systematic model errors or biases, AVII
227
Annex VII Glossary
which may be assessed from systematic comparisons of model tolerance, CIDs and their changes can be detrimental, beneficial,
simulations with observations wherever available. neutral or a mixture of each across interacting system elements and
regions. See also Risk, Hazard and Impacts (consequences, outcomes).
Climate system The global system consisting of five major
components: the atmosphere, the hydrosphere, the cryosphere, the Cloud condensation nuclei (CCN) The subset of aerosol
lithosphere and the biosphere and the interactions between them. particles that serve as an initial site for the condensation of liquid
The climate system changes in time under the influence of its own water, which can lead to the formation of cloud droplets, under typical
internal dynamics and because of external forcings such as volcanic cloud formation conditions. The main factor that determines which
eruptions, solar variations, orbital forcing, and anthropogenic aerosol particles are CCN at a given supersaturation is their size.
forcings such as the changing composition of the atmosphere and
Cloud feedback A climate feedback involving changes in any
land-use change.
of the properties of clouds as a response to a change in the local
Climate threshold A limit within the climate system (or its or global surface temperature. Understanding cloud feedbacks and
forcing) beyond which the behaviour of the system is qualitatively determining their magnitude and sign requires an understanding of
changed. See also Abrupt climate change and Tipping point. how a change in climate may affect the spectrum of cloud types,
the cloud fraction and height, the radiative properties of clouds, and
Climate variability Deviations of climate variables from
finally the Earth’s radiation budget.
a given mean state (including the occurrence of extremes, etc.) at
all spatial and temporal scales beyond that of individual weather Cloud radiative effect The radiative effect of clouds relative to
events. Variability may be intrinsic, due to fluctuations of processes the identical situation without clouds.
internal to the climate system (internal variability), or extrinsic, due
Cloud-resolving models (CRMs) Numerical models that are
to variations in natural or anthropogenic external forcing (forced
that are of high enough resolution and have the necessary physics to
variability). See also Climate change and Modes of climate variability.
represent the dynamical and physical processes of cloud formation.
Decadal variability
CMIP6 See Coupled Model Intercomparison Project (CMIP).
Decadal variability refers to climate variability on decadal time scales.
See also Pacific Decadal Variability (PDV), Atlantic Multi-decadal CO2 equivalent (CO2-eq) emission The amount of carbon
Oscillation/Variability (AMO/AMV) and Pacific Decadal Oscillation dioxide (CO2) emission that would have an equivalent effect on
(PDO) (under Pacific Decadal Variability (PDV)). a specified key measure of climate change, over a specified time
horizon, as an emitted amount of another greenhouse gas (GHG) or
Internal variability
a mixture of other GHGs. For a mix of GHGs, it is obtained by summing
Fluctuations of the climate dynamical system when subject to
the CO2-equivalent emissions of each gas. There are various ways and
a constant or periodic external forcing (such as the annual cycle). See
time horizons to compute such equivalent emissions (see greenhouse
also Climate variability.
gas emission metric). CO2-equivalent emissions are commonly used
Natural variability to compare emissions of different GHGs, but should not be taken
Natural variability refers to climatic fluctuations that occur without to imply that these emissions have an equivalent effect across all
any human influence, that is, internal variability combined with key measures of climate change. [Note: Under the Paris Rulebook
the response to external natural factors such as volcanic eruptions, (Decision 18/CMA.1, annex, paragraph 37), parties have agreed to
changes in solar activity and, on longer time scales, orbital effects use GWP-100 values from the IPCC AR5 or GWP-100 values from a
and plate tectonics. See also Orbital forcing. subsequent IPCC Assessment Report to report aggregate emissions
and removals of GHGs. In addition, parties may use other metrics
Climate velocity The speed at which isolines of a specified
to report supplemental information on aggregate emissions and
climate variable travel across landscapes or seascapes due to
removals of GHGs.]
changing climate. For example, climate velocity for temperature is the
speed at which isotherms move due to changing climate (km yr–1) and Coast The land near to the sea. The term ‘coastal’ can refer to that
is calculated as the temporal change in temperature (°C yr–1) divided land (e.g., as in ‘coastal communities’), or to that part of the marine
by the current spatial gradient in temperature (°C km–1). It can be environment that is strongly influenced by land-based processes. Thus,
calculated using additional climate variables such as precipitation or coastal seas are generally shallow and near-shore. The landward and
can be based on the climatic niche of organisms. seaward limits of the coastal zone are not consistently defined, either
scientifically or legally. Thus, coastal waters can either be considered
Climate–carbon cycle feedback A climate feedback involves
as equivalent to territorial waters (extending 12 nautical miles/22.2
changes in the properties of the land and ocean carbon cycle in
km from mean low water), or to the full Exclusive Economic Zone, or
response to climate change. In the ocean, changes in oceanic
to shelf seas, with less than 200 m water depth.
temperature and circulation could affect the atmosphere–ocean
carbon dioxide (CO2) flux; on the continents, climate change could Common era (CE) CE (Common Era) and BCE (Before the
affect plant photosynthesis and soil microbial respiration and hence Common Era) are alternative names for AD (Anno Domini) and BC
the flux of CO2 between the atmosphere and the land biosphere. (Before Christ) in the Gregorian international standard calendar-year
system. CE/BCE are preferred in an international context because
Climatic impact-driver (CID) Climatic impact-drivers (CIDs) are
they are neutral with respect to religion. The numbering of calendar
AVII physical climate system conditions (e.g., means, events, extremes)
that affect an element of society or ecosystems. Depending on system
228
Glossary Annex VII
years is the same under both terminologies. The CE began in year based on shared model inputs by modelling groups from around
AD 1 and extends to the present day. the world. The CMIP Phase 3 (CMIP3) multi-model dataset includes
projections using Special Report on Emissions Scenarios (SRES)
Compatible emissions Earth system models that simulate
scenarios. The CMIP Phase 5 (CMIP5) dataset includes projections
the land and ocean carbon cycle can calculate carbon dioxide
using the Representative Concentration Pathways (RCP). The CMIP6
(CO2) emissions that are compatible with a given atmospheric CO2
phase involves a suite of common model experiments as well as an
concentration trajectory. The compatible emissions over a given period
ensemble of CMIP-endorsed Model Intercomparison Projects (MIPs).
of time are equal to the increase of carbon over that same period of
time in the sum of the three active reservoirs: the atmosphere, the Cryosphere The components of the Earth system at and below
land and the ocean. the land and ocean surface that are frozen, including snow cover,
glaciers, ice sheets, ice shelves, icebergs, sea ice, lake ice, river ice,
Compound events See Compound weather/climate events.
permafrost and seasonally frozen ground.
Compound weather/climate events The terms ‘compound
Cumulative emissions The total amount of emissions released
events’, ‘compound extremes’ and ‘compound extreme events’ are
over a specified period of time. See also Carbon budget and Transient
used interchangeably in the literature and this report and refer to the
climate response to cumulative CO2 emissions (TCRE) (under Climate
combination of multiple drivers and/or hazards that contributes to
sensitivity).
societal and/or environmental risk (Zscheischler et al., 2018).
Dansgaard–Oeschger events (D-O events) Millennial-scale
Concentrations scenario See Scenario.
events first characterized in Greenland ice cores as abrupt warming
Confidence The robustness of a finding based on the type, from a cold stadial state to a warmer interstadial state, followed by
amount, quality and consistency of evidence (e.g., mechanistic a return to a cold stadial state (Dansgaard et al., 1993), and traced in
understanding, theory, data, models, expert judgement) and on the the ocean via deposits of ice-rafted sand grains (Bond and Lotti, 1995).
degree of agreement across multiple lines of evidence. In this report, Named after Willi Dansgaard and Hans Oeschger by Bond and Lotti
confidence is expressed qualitatively (Mastrandrea et al., 2010). (1995). An example of a D–O event during the most recent deglacial
transition is the Bølling–Allerød interstadial. Warm D–O events in
Constant composition commitment See Climate change
Greenland are associated with cooling events in Antarctica (Blunier
commitment.
and Brook, 2001) through ocean thermohaline circulation (Stocker
Constant emissions commitment See Climate change and Johnsen, 2003). See also Bipolar seesaw (also interhemispheric
commitment. seesaw, interhemispheric asymmetry, hemispheric asymmetry).
Convection Vertical motion driven by buoyancy forces arising Data assimilation Mathematical method used to combine
from static instability, usually caused by near-surface cooling or different sources of information in order to produce the best possible
increases in salinity in the case of the ocean and near-surface estimate of the state of a system. This information usually consists
warming or cloud-top radiative cooling in the case of the atmosphere. of observations of the system and a numerical model of the system
In the atmosphere, convection gives rise to cumulus clouds and evolution. Data assimilation techniques are used to create initial
precipitation and is effective at both scavenging and vertically conditions for weather forecast models and to construct reanalyses
transporting chemical species. In the ocean, convection can carry describing the trajectory of the climate system over the time period
surface waters to deep within the ocean. covered by the observations.
Convection-permitting models See Cloud-resolving models Dead zones Extremely hypoxic (i.e., low-oxygen) areas in oceans
(CRMs). and lakes, caused by excessive nutrient input from human activities
coupled with other factors that deplete the oxygen required to
Coral bleaching Loss of coral pigmentation through the loss of support many marine organisms in bottom and near-bottom water.
intracellular symbiotic algae (known as zooxanthellae) and/or loss
of their pigments. Decadal predictability Refers to the notion of predictability
of the climate system on a decadal time scale. See also Climate
Coral reef An underwater ecosystem characterised by structure- prediction, Predictability and Decadal prediction.
building stony corals. Warm-water coral reefs occur in shallow seas,
mostly in the tropics, with the corals (animals) containing algae Decadal prediction A climate prediction on decadal time scales.
(plants) that depend on light and relatively stable temperature See also Predictability and Decadal predictability.
conditions. Cold-water coral reefs occur throughout the world,
Decadal variability See Climate variability.
mostly at water depths of 50–500 m. In both kinds of reef, living
corals frequently grow on older, dead material, predominantly made Deep uncertainty See Uncertainty.
of calcium carbonate (CaCO3). Both warm- and cold-water coral reefs
Deforestation Conversion of forest to non-forest. [Note: For
support high biodiversity of fish and other groups and are considered
a discussion of the term forest and related terms such as afforestation,
to be especially vulnerable to climate change.
reforestation and deforestation, see the 2006 IPCC Guidelines for
Coupled Model Intercomparison Project (CMIP) A climate National Greenhouse Gas Inventories and their 2019 Refinement, and
modelling activity from the World Climate Research Programme information provided by the United Nations Framework Convention
AVII
(WCRP), which coordinates and archives climate model simulations
229
Annex VII Glossary
on Climate Change (IPCC, 2006, 2019; UNFCCC, 2021a, b).] See also Dobson unit (DU) A unit to measure the total amount of ozone
Afforestation and Reforestation. in a vertical column above the Earth’s surface (total column ozone).
The number of Dobson units is the thickness in units of 10–5 m that
Deglacial or deglaciation or glacial termination The period
the ozone column would occupy if compressed into a layer of uniform
of transition from glacial conditions at the end of a glacial period
density at a pressure of 1013 hPa and a temperature of 0°C. One DU
to interglacial conditions characterized by a reduction in land ice
corresponds to a column of ozone containing 2.69 × 1020 molecules
volume. Gradual changes can be punctuated by abrupt changes
per square metre. A typical value for the amount of ozone in a column
linked to stadial/interstadial events and bipolar seesaw aspect. The
of the Earth’s atmosphere, although very variable, is 300 DU.
last deglacial transition occurred between about 18,000 and 11,000
years ago. It encompasses rapid events such as Meltwater Pulse 1A Downscaling A method that derives local- to regional-scale
(MWP-1A) and millennial-scale fluctuations such as the Younger information from larger-scale models or data analyses. Two main
Dryas. See also Glacial–interglacial cycles and Ice age. methods exist: dynamical downscaling and empirical/statistical
downscaling. The dynamical method uses the output of regional
Detection Detection of change is defined as the process of
climate models, global models with variable spatial resolution, or
demonstrating that climate or a system affected by climate has
high-resolution global models. The empirical/statistical methods are
changed in some defined statistical sense, without providing a reason
based on observations and develop statistical relationships that link
for that change. An identified change is detected in observations if its
the large-scale atmospheric variables with local/regional climate
likelihood of occurrence by chance due to internal variability alone is
variables. In all cases, the quality of the driving model remains an
determined to be small, for example, <10%.
important limitation on quality of the downscaled information. The
Detection and attribution See Detection and Attribution. two methods can be combined, for example, applying empirical/
statistical downscaling to the output of a regional climate model
Diatoms Microscopic (2–200 µm) unicellular photosynthetic
consisting of a dynamical downscaling of a global climate model.
algae that live in surface waters of lakes, rivers and oceans and form
shells of opal. In the global ocean, marine diatom species distribution Drought An exceptional period of water shortage for existing
is primarily driven by nutrient availability. On regional scales, their ecosystems and the human population (due to low rainfall, high
species distribution in ocean sediment cores can be related to past temperature, and/or wind). See also Plant evaporative stress.
sea surface temperatures (Abrantes et al., 2013).
Agricultural and ecological drought
Dimensions of integration In IPCC AR6, concepts used to Depending on the affected biome: a period with abnormal soil moisture
synthesize the knowledge of climate change across not just the deficit, which results from combined shortage of precipitation and
physical sciences, but also across impacts, adaptation and mitigation excess evapotranspiration, and during the growing season impinges
research. The concept of ‘dimensions of integration’ includes on crop production or ecosystem function in general.
(i) emission and concentration scenarios underlying the climate
Hydrological drought
change projections assessed in this report, (ii) levels of projected
A period with large runoff and water deficits in rivers, lakes and
global mean temperature change and (iii) total amounts of cumulative
reservoirs.
carbon emissions for projections.
Meteorological drought
Direct (aerosol) effect See Aerosol–radiation interaction.
A period with an abnormal precipitation deficit.
Direct air capture (DAC) Chemical process by which a pure
Dynamic global vegetation model (DGVM) A model that
carbon dioxide (CO2) stream is produced by capturing CO2 from the
simulates vegetation development and dynamics through space and
ambient air. See also Anthropogenic removals and Carbon dioxide
time, as driven by climate and other environmental changes.
removal (CDR).
Dynamical downscaling See Downscaling.
Disaster A ‘serious disruption of the functioning of a community
or a society at any scale due to hazardous events interacting with Dynamical system A process or set of processes whose evolution
conditions of exposure, vulnerability and capacity, leading to one or in time is governed by a set of deterministic physical laws. The climate
more of the following: human, material, economic and environmental system is a dynamical system.
losses and impacts’ (UNGA, 2016). See also Exposure, Hazard, Risk
Early Eocene Climatic Optimum (EECO) The EECO is a period
and Vulnerability.
of geological time that occurred about 53 to 49 million years ago,
Discharge (of ice) See Mass balance/budget (of glaciers or ice during the Eocene Epoch. Continental positions at this time were
sheets). somewhat different to present due to tectonic plate movements.
Geological data indicate that the EECO was a period of relatively
Dissolved inorganic carbon The combined total of different
high atmospheric CO2 concentrations (about 1150–2500 ppmv)
types of non-organic carbon in (seawater) solution, comprising
and relative warmth (global mean surface temperature was about
carbonate (CO32–), bicarbonate (HCO3–), carbonic acid (H2CO3) and
10–18°C above the 1850–1900 reference), and polar ice sheets
carbon dioxide (CO2).
were absent.
Diurnal temperature range (DTR) The difference between the
Earth system model (ESM) A coupled atmosphere–ocean
AVII maximum and minimum temperature during a 24-hour period.
general circulation model (AOGCM) in which a representation of
230
Glossary Annex VII
the carbon cycle is included, allowing for interactive calculation Ecosystem A functional unit consisting of living organisms, their
of atmospheric carbon dioxide (CO2) or compatible emissions. non-living environment and the interactions within and between
Additional components (e.g., atmospheric chemistry, ice sheets, them. The components included in a given ecosystem and its spatial
dynamic vegetation, nitrogen cycle, but also urban or crop models) boundaries depend on the purpose for which the ecosystem is
may be included. See also Earth system model of intermediate defined: in some cases, they are relatively sharp, while in others they
complexity (EMIC). are diffuse. Ecosystem boundaries can change over time. Ecosystems
are nested within other ecosystems, and their scale can range from
Earth system model of intermediate complexity
very small to the entire biosphere. In the current era, most ecosystems
(EMIC) EMICs represent climate processes at a lower resolution
either contain people as key organisms or are influenced by the
or in a simpler, more idealized fashion than an Earth system model
effects of human activities in their environment.
(ESM).
Effective radiative forcing (ERF) See Radiative forcing,
Earth’s energy budget encompasses the major energy flows
Aerosol effective radiative forcing (ERFari+aci) (under Aerosol–
of relevance for the climate system: the top-of-atmosphere energy
radiation interaction), Effective radiative forcing (or effect) due to
budget; the surface energy budget; changes in the global energy
aerosol–cloud interactions (ERFaci) (under Aerosol–cloud interaction)
inventory and internal flows of energy within the climate system that
and Effective radiative forcing (or effect) due to aerosol–radiation
characterize the climate state.
interactions (ERFari) (under Aerosol–radiation interaction).
Top-of-atmosphere energy budget
Ekman transport The total transport resulting from a balance
Comprises the energy fluxes associated with incoming solar radiation,
between the Coriolis force and the frictional stress due to the action
reflected solar radiation and emitted thermal radiation. Typical units:
of the wind on the ocean surface.
W m–2.
El Niño See El Niño–Southern Oscillation (ENSO).
Surface energy budget
Comprises the exchanges of heat at the surface of the Earth associated El Niño–Southern Oscillation (ENSO) The term El Niño was
with both radiative and non-radiative processes. Typical units: W m–2. initially used to describe a warm-water current that periodically flows
along the coast of Ecuador and Peru, disrupting the local fishery. It has
Global energy inventory
since become identified with warming of the tropical Pacific Ocean
Quantifies the excess energy absorbed or lost by the Earth system
east of the dateline. This oceanic event is associated with a fluctuation
(ocean, land, atmosphere and cryosphere), mostly in the form of heat,
of a global-scale tropical and subtropical surface pressure pattern
associated with radiative forcing of the climate. Typical units: Joules.
called the Southern Oscillation. This coupled atmosphere–ocean
Global energy budget phenomenon, with preferred time scales of two to about seven years,
For a given time period, the global energy budget expresses the is known as the El Niño–Southern Oscillation (ENSO). The warm and
balance between change in the global energy inventory, the time- cold phases of ENSO are called El Niño and La Niña, respectively. ENSO
integrated effective radiative forcing and time-integrated radiative is often measured by the surface pressure anomaly difference between
response of the climate system. Typical units: Joules. Tahiti and Darwin and/or the sea surface temperatures in the central
and eastern equatorial Pacific. This phenomenon has a great impact
See also Earth’s energy imbalance.
on the wind, sea surface temperature and precipitation patterns in the
Earth’s energy imbalance The persistent and positive tropical Pacific. It has climatic effects throughout the Pacific region
(downward) net top of atmosphere energy flux associated with and in many other parts of the world through global teleconnections.
greenhouse gas forcing of the climate system. See also Earth’s energy See Section AIV.2.3 in Annex IV of the AR6 WGI report.
budget and Radiative response (of the climate system).
Central Pacific El Niño
Earth system sensitivity See Climate sensitivity. An El Niño event in which sea surface temperature anomalies are
stronger in the central equatorial Pacific than in the east. Also known
Effective equilibrium climate sensitivity See Climate as a Modoki El Niño event.
sensitivity.
Eastern Pacific El Niño
East Antarctic Ice Sheet (EAIS) See Ice sheet. An El Niño event in which sea surface temperature anomalies are
East Asian monsoon (EAsiaM) See Global monsoon. largest in the eastern tropical Pacific.
Eastern boundary upwelling systems (EBUS) Eastern Electromagnetic spectrum Wavelength, frequency or energy
boundary upwelling systems (EBUS) are located at the eastern range of all electromagnetic radiation. In terms of solar radiation, the
(landward) edges of major ocean basins in both hemispheres, where spectral irradiance is the power arriving at the Earth per unit area,
equatorward winds drive upwelling currents that bring cool, nutrient- per unit wavelength.
rich (and often oxygen-poor) waters from the deep ocean to the Elevation-dependent warming (EDW) Characteristic of many
surface near the coast. regions where mountains are located, in which past and/or future
Eastern Pacific El Niño See El Niño–Southern Oscillation (ENSO). surface air temperature changes vary neither uniformly nor linearly
with elevation. In many cases, warming is enhanced within or above
Economic potential See Mitigation potential. AVII
a certain elevation range.
231
Annex VII Glossary
Emergence (of the climate signal) Emergence of a climate Energy balance The difference between the total incoming and
change signal or trend refers to when a change in climate (the ‘signal’) total outgoing energy. If this balance is positive, warming occurs; if
becomes larger than the amplitude of natural or internal variations it is negative, cooling occurs. Averaged over the globe and over long
(defining the ‘noise’), This concept is often expressed as a ‘signal- time periods, this balance must be zero. Because the climate system
to-noise’ ratio and emergence occurs at a defined threshold of this derives virtually all its energy from the Sun, zero balance implies that,
ratio (e.g., S/N > 1 or 2). Emergence can refer to changes relative to globally, the absorbed solar radiation, that is, incoming solar radiation
a historical or modern baseline (usually at least 20 years long) and minus reflected solar radiation at the top of the atmosphere and
can also be expressed in terms of time (time of emergence) or in terms outgoing longwave radiation emitted by the climate system are equal.
of a global warming level. Emergence is also used to refer to a time
Energy budget (of the Earth) The Earth is a physical system
when we can expect to see a response to reducing greenhouse gas
with an energy budget that includes all gains of incoming energy and
(GHG) emissions (emergence with respect to mitigation). Emergence
all losses of outgoing energy. The Earth’s energy budget is determined
can be estimated using observations and/or model simulations. See
by measuring how much energy comes into the Earth system from
also Time of emergence (ToE).
the Sun, how much energy is lost to space, and accounting for
Emergent constraint An attempt to reduce the uncertainty in the remainder on Earth and its atmosphere. Solar radiation is the
climate projections, using an ensemble of Earth system models (ESMs) dominant source of energy into the Earth system. Incoming solar
to relate a specific feedback or future change to an observation of the energy may be scattered and reflected by clouds and aerosols or
past or current climate (typically some trend, variability or change absorbed in the atmosphere. The transmitted radiation is then either
in variability). absorbed or reflected at the Earth’s surface. The average albedo of
the Earth is about 0.3, which means that 30% of the incident solar
Emission factor/Emissions intensity A coefficient that
energy is reflected into space, while 70% is absorbed by the Earth.
quantifies the emissions or removals of a gas per unit activity.
Radiant solar or shortwave energy is transformed into sensible heat,
Emission factors are often based on a sample of measurement data,
latent energy (involving different water states), potential energy, and
averaged to develop a representative rate of emission for a given
kinetic energy before being emitted as infrared radiation. With the
activity level under a given set of operating conditions.
average surface temperature of the Earth of about 15°C (288 K),
Emission pathways See Pathways. the main outgoing energy flux is in the infrared part of the spectrum.
See also Sensible heat flux and Latent heat flux.
Emissions See Cumulative emissions, Anthropogenic emissions,
Fossil fuel emissions, Non-CO2 emissions and radiative forcing and Enhanced weathering A proposed method to increase the
Negative greenhouse gas emissions. See also Emissions scenario natural rate of removal of carbon dioxide (CO2) from the atmosphere
(under Scenario), and Emission pathways. using silicate and carbonate rocks. The active surface area of these
minerals is increased by grinding, before they are actively added to
Emulation Reproducing the behaviour of complex, process-
soil, beaches or the open ocean. See also Carbon dioxide removal
based models (namely, Earth system models, ESMs) via simpler
(CDR) and Anthropogenic removals.
approaches, using either emulators or simple climate models (SCMs).
The computational efficiency of emulating approaches opens new Ensemble A collection of comparable datasets that reflect
analytical possibilities given that ESMs take a lot of computational variations within the bounds of one or more sources of uncertainty
resources for each simulation. See also Emulators and Simple climate and that, when averaged, can provide a more robust estimate
model (SCM). of underlying behaviour. Ensemble techniques are used by the
observational, reanalysis and modelling communities. See also
Emulators A broad class of heavily parametrized models (‘simple
Climate simulation ensemble.
climate models’), statistical methods like neural networks, genetic
algorithms or other artificial intelligence approaches designed to Equilibrium and transient climate experiment An
reproduce the responses of more complex, process-based Earth equilibrium climate experiment is a climate model experiment in
system models (ESMs). The main application of emulators is to which the model is allowed to fully adjust to a change in radiative
extrapolate insights from ESMs and observational constraints forcing. Such experiments provide information on the difference
to a larger set of emission scenarios. See also Emulation and Simple between the initial and final states of the model, but not on the time-
climate model (SCM). dependent response. If the forcing is allowed to evolve gradually
according to a prescribed emissions scenario, the time-dependent
Energy balance model (EBM) An energy balance model is
response of a climate model may be analysed. Such an experiment is
a simplified climate model that is typically used as an emulator
called a transient climate experiment.
of climate to analyse the energy budget of the Earth to compute
changes in the climate. In its simplest form, there is no explicit spatial Equilibrium climate sensitivity (ECS) See Climate sensitivity.
dimension, and the model then provides an estimate of the changes
Equilibrium line The spatially averaged boundary at a given
in globally averaged temperature computed from the changes
moment, usually chosen as the seasonal mass budget minimum at
in radiation. This zero-dimensional energy balance model can be
the end of summer, between the region on a glacier where there
extended to a one-dimensional or two-dimensional model if changes
is a net annual loss of ice mass (ablation area) and that where there is
to the energy budget with respect to latitude, or both latitude and
a net annual gain (accumulation area). The altitude of this boundary
AVII longitude, are explicitly considered.
is referred to as equilibrium line altitude (ELA).
232
Glossary Annex VII
Equivalent carbon dioxide (CO2) emission See CO2 equivalent modification of coastlines or dredging. In turn, changes in any or all
(CO2-eq) emission. of the contributions to extreme sea levels may lead to long–term
relative sea level changes. Alternate expressions for ESL may be used
Eutrophication Over-enrichment of water by nutrients such as
depending on the processes resolved.
nitrogen and phosphorus. It is one of the leading causes of water
quality impairment. The two most acute symptoms of eutrophication Extreme still water level (ESWL) refers to the combined contribution
are hypoxia (or oxygen depletion) and harmful algal blooms. of relative sea level change, tides and storm-surges. Wind-waves
also contribute to coastal sea level via three processes: infragravity
Evaporation The physical process by which a liquid (e.g., water)
waves (lower frequency gravity waves generated by the wind waves),
becomes a gas (e.g., water vapour).
wave setup (time-mean sea level elevation due to wave energy
Evapotranspiration The combined processes through which dissipation), and swash (vertical displacement up the shore-face
water is transferred to the atmosphere from open water and ice induced by individual waves). Extreme total water level (ETWL) is the
surfaces, bare soil, and vegetation that make up the Earth’s surface. ESWL plus wave setup. When considering coastal impacts, swash is
also important, and Extreme coastal water level (ECWL) is used. See
Potential evapotranspiration The potential rate of water loss from
also Storm surge and Sea level change (sea level rise/sea level fall).
wet soils and from plant surfaces, without any limits imposed by the
water supply. Extreme still water level (ESWL) See Extreme sea level (ESL).
Evidence Data and information used in the scientific process to Extreme total water level (ETWL) See Extreme sea level (ESL).
establish findings. In this report, the degree of evidence reflects the
Extreme weather event An event that is rare at a particular
amount, quality and consistency of scientific/technical information
place and time of year. Definitions of ‘rare’ vary, but an extreme
on which the Lead Authors are basing their findings. See also
weather event would normally be as rare as or rarer than the 10th
Agreement, Confidence, Likelihood and Uncertainty.
or 90th percentile of a probability density function estimated from
Exposure The presence of people; livelihoods; species or observations. By definition, the characteristics of what is called
ecosystems; environmental functions, services, and resources; extreme weather may vary from place to place in an absolute sense.
infrastructure; or economic, social, or cultural assets in places and See also Climate extreme (extreme weather or climate event).
settings that could be adversely affected.
Extreme/heavy precipitation event An extreme/heavy
Extended concentration pathways (ECPs) See Representative precipitation event is an event that is of very high magnitude with a
concentration pathways (RCPs) (under Pathways). very rare occurrence at a particular place. Types of extreme precipitation
may vary depending on its duration (hourly, daily or multi-days (e.g., 5
External forcing External forcing refers to a forcing agent
days)) though all of them qualitatively represent high magnitude. The
outside the climate system causing a change in the climate system.
intensity of such events may be defined with a block maxima approach
Volcanic eruptions, solar variations and changes in Earth’s orbit, as
such as annual maxima or with a peaks over threshold approach, such
well as anthropogenic changes in the composition of the atmosphere
as rainfall above the 95th or 99th percentile at a particular place.
or in land use are external forcings. See also Orbital forcing.
Faculae Bright patches on the Sun. The area covered by faculae is
Extratropical cyclone (ETC) Any cyclonic-scale storm that is not
greater during periods of high solar activity.
a tropical cyclone. Usually refers to a mid- or high-latitude migratory
storm system formed in regions of large horizontal temperature Feedback See Climate feedback.
variations. Sometimes called extratropical storm or extratropical low.
Fine-mode aerosol optical depth See Aerosol optical depth
Extratropical jets Extratropical jets are wind maxima in the (AOD).
upper troposphere marking zones of baroclinic instability. Anomalies
Fingerprint The climate response pattern in space and/or time to
in the position of these jets are often associated with storms, blocking,
a specific forcing is commonly referred to as a fingerprint. The spatial
and weather extremes.
patterns of sea level response to melting of glaciers or ice sheets (or
Extreme climate event See Climate extreme (extreme weather other changes in surface loading) are also referred to as fingerprints.
or climate event). Fingerprints are used to detect the presence of this response in
observations and are typically estimated using forced climate model
Extreme coastal water level (ECWL) See Extreme sea
simulations. See also Detection and attribution.
level (ESL).
Fire weather Weather conditions conducive to triggering
Extreme sea level (ESL) The occurrence of an exceptionally
and sustaining wildfires, usually based on a set of indicators and
low or high local sea surface height, arising from (a combination
combinations of indicators including temperature, soil moisture,
of) short-term phenomena (e.g., storm surges, tides and waves).
humidity, and wind. Fire weather does not include the presence or
Relative sea level changes affect extreme sea levels directly by
absence of fuel load.
shifting the mean water levels and indirectly by modulating the
propagation of tides, waves and/or surges due to increased water Firn Snow that has survived at least one ablation season but has
depth. In addition, extreme sea levels can be influenced by changes not been transformed to glacier ice. Its pore space is at least partially
in the frequency, tracks or strength of weather systems and interconnected, allowing air and water to circulate. Firn densities AVII
storms, or due to anthropogenically induced changes such as the typically are 400–830 kg m–3.
233
Annex VII Glossary
Fitness-for-purpose The suitability of a model (or other resource, General circulation model (GCM) A numerical representation
such as a dataset or method) for a particular task, such as quantifying of the atmosphere–ocean–sea ice system based on the physical,
the contribution of increased greenhouse gas concentrations to chemical and biological properties of its components, their
recent changes in global mean surface temperature or projecting interactions and feedback processes. General circulation models
changes in drought frequency in a region under a given scenario. are used for weather forecasts, seasonal to decadal prediction, and
Assessment of a model’s fitness-for-purpose can be informed both by climate projections. They are the basis of the more complex Earth
how the model represents relevant physical processes and by how it system models (ESMs). See also Climate model.
scores on relevant performance metrics.
Geocentric sea level change See Sea level change (sea level
Flaring Open air burning of waste gases and volatile liquids, rise/sea level fall).
through a chimney, at oil wells or rigs, in refineries or chemical plants,
Geoid The equipotential surface having the same geopotential at
and at landfills.
each latitude and longitude around the world (geodesists denote this
Flood The overflowing of the normal confines of a stream or other potential W0) that best approximates the mean sea level. It is the
water body, or the accumulation of water over areas that are not surface of reference for measurement of altitude. In practice, several
normally submerged. Floods can be caused by unusually heavy rain, variations of definitions of the geoid exist depending on the way the
for example during storms and cyclones. Floods include river (fluvial) permanent tide (the zero-frequency gravitational tide due to the Sun
floods, flash floods, urban floods, rain (pluvial) floods, sewer floods, and Moon) is considered in geodetic studies.
coastal floods and glacial lake outburst floods (GLOFs).
Geostrophic winds or currents A wind or current that is in
Flux A movement (a flow) of matter (e.g., water vapour, particles), balance with the horizontal pressure gradient and the Coriolis force,
heat or energy from one place to another, or from one medium (e.g., and thus is outside of the influence of friction. Thus, the wind or
land surface) to another (e.g., atmosphere). current is directly parallel to isobars and its speed is proportional to
the horizontal pressure gradient.
Foraminifera Single-celled, sand-sized marine organisms
(protists) that possess a hard test mainly composed of agglutinated Glacial isostatic adjustment (GIA) The ongoing changes
walls (detrital grains glued together with organic cement) or calcium in gravity, rotation and viscoelastic solid Earth deformation (GRD)
carbonate (predominantly calcite). They are used to reconstruct in response to past changes in the distribution of ice and water
a range of (paleo)environmental variables such as salinity, on Earth’s surface. On a time scale of decades to tens of millennia
temperature, oxygenation, oxygen isotope composition and organic following mass redistribution, Earth’s mantle flows viscously as it
and nutrient flux. evolves toward isostatic equilibrium, causing solid Earth movement
and geoid changes, which can result in regional-to-local sea level
Forcing See Radiative forcing.
variations. See also Sea level change (sea level rise/sea level fall).
Forest A vegetation type dominated by trees. Many definitions
Glacial lake outburst flood (GLOF)/Glacier lake
of the term forest are in use throughout the world, reflecting
outburst A sudden release of water from a glacier lake, including
wide differences in biogeophysical conditions, social structure
any of the following types – a glacier-dammed lake, a pro-glacial
and economics. [Note: For a discussion of the term forest in the
moraine-dammed lake or water that was stored within, under or on
context of National GHG inventories, see the 2006 IPCC Guidelines
the glacier.
for National GHG Inventories and their 2019 Refinement, and
information provided by the United Nations Framework Convention Glacial or glaciation A period characterized by the establishment
on Climate Change (IPCC, 2006, 2019; UNFCCC, 2021a, b).] See also of expanded ice sheets and glaciers, and associated with global mean
Afforestation, Deforestation and Reforestation. sea level (GMSL) substantially lower than present; generally coincides
with even-numbered marine isotope stages. Glacial intervals were
Fossil fuel emissions Emissions of greenhouse gases (GHGs)
interrupted by interglacial intervals. The Last Glacial Maximum (LGM)
(in particular carbon dioxide (CO2)), other trace gases and aerosols
is a specific interval within the most recent glaciation, when ice sheets
resulting from the combustion of fuels from fossil carbon deposits
were near their global maximum volume (Clark et al., 2009; Gowan
such as oil, gas and coal.
et al., 2021) and GMSL was nearly at its lowest level (Lambeck et al.,
Fossil fuels Carbon-based fuels from fossil hydrocarbon deposits, 2014; Yokoyama et al., 2018). Local or regional glacial maxima may
including coal, oil and natural gas. be diachronous, for example ranging from about 29,000 years ago
and 16,000 years ago. For purposes of global synthesis, IPCC AR6
Free atmosphere The atmospheric layer that is negligibly
adopts a practical chronostratigraphic definition of LGM of 23,000–
affected by friction against the Earth’s surface, and which is above
19,000 years BP (before 1950; chronozone level 1 of Mix et al., 2001).
the atmospheric boundary layer.
For modelling purposes, LGM is defined by the model time step
Frozen ground Soil or rock in which part or all of the pore water nearest to the centre of this interval, 21,000 years ago (Kageyama
consists of ice. See also Active layer and Permafrost. et al., 2017). See also Deglacial or deglaciation or glacial termination,
Glacial–interglacial cycles, Ice age and Interglacial or interglaciation.
General circulation The large-scale motions of the atmosphere
and the ocean as a consequence of differential heating on a rotating Glacial termination See Deglacial or deglaciation or glacial
AVII Earth. General circulation contributes to the energy balance of the termination.
system through transport of heat and momentum.
234
Glossary Annex VII
Glacial–interglacial cycles Phase of the Earth’s history marked convection and heavy rainfall. Over northern Australia, the monsoon
by large changes in continental ice volume and global sea level. season generally lasts from December to March and is associated
See also Glacial or glaciation, Deglacial or deglaciation or glacial with west to north-westerly inflow of moist winds, producing
termination, Interglacial or interglaciation and Ice age. convection and heavy precipitation. Over the Maritime Continent,
the main rainy season south of the equator is centred on December
Glaciated State of a surface that was covered by glacier ice in the
to February with north-westerly monsoon flow at low levels. Further
past, but not at present. See also Glacierized.
details on how AusMCM is defined and used throughout the Report
Glacier A perennial mass of ice, and possibly firn and snow, are provided in Annex V.
originating on the land surface by accumulation and compaction of
East Asian monsoon (EAsiaM)
snow and showing evidence of past or present flow. A glacier typically
The East Asian monsoon (EAsiaM) is the seasonal reversal in wind
gains mass by accumulation of snow and loses mass by ablation.
and precipitation occurring over East Asia, including eastern China,
Land ice masses of continental size (>50,000 km2) are referred to as
Japan and the Korean peninsula. In contrast to the other monsoons
ice sheets (Cogley et al., 2011).
it extends quite far north, out of the tropical belt, and it is largely
Outlet glacier A glacier, usually between rock walls, that is part of, influenced by subtropical systems and by disturbances from the mid-
and drains, an ice sheet. See also Ice stream. latitudes. The EAsiaM manifests during boreal summer with warm
and wet southerly winds, but also during boreal winter with cold
Glacierized A surface that is currently covered by glacier ice. See
and dry northerly winds. In late April/early May, rainfall onsets in
also Glaciated.
the central Indochina Peninsula, and in mid-June the rainy season
Global carbon budget See Carbon budget. arrives over East Asia with the formation of the Meiyu front along
the Yangtze River valley, Changma in Korea and Baiu in Japan. In
Global dimming Global dimming refers to the observed July, the monsoon advances up to North China, the Korean peninsula
widespread reduction in the amount of solar radiation received at and central Japan. During boreal winter, strong north-westerlies
the Earth’s surface from the 1950s to the 1980s, with an increase in manifest over north and north-east China, Korea and Japan, while
anthropogenic aerosol emissions appearing to have contributed. This strong north-easterlies arrive along the coast of East Asia. Further
was followed by a partial recovery since the 1990s (‘brightening’), details on how EAsiaM is defined and used throughout the Report
particularly in industrialized areas, coincident with a reduction in are provided in Annex V.
anthropogenic aerosol emissions.
North American monsoon (NAmerM)
Global mean sea level (GMSL) change See Sea level change The North American monsoon (NAmerM) is a regional-scale
(sea level rise/sea level fall). atmospheric circulation system with increases in summer
Global mean surface air temperature (GSAT) Global precipitation over northwestern Mexico and southwest United States.
average of near-surface air temperatures over land, oceans and The monsoonal characteristics of the region include a pronounced
sea ice. Changes in GSAT are often used as a measure of global annual maximum of precipitation in boreal summer (June–July–
temperature change in climate models. See also Global mean surface August) accompanied by a surface low pressure system and an
temperature (GMST). upper-level anticyclone, although seasonal reversal of the surface
winds is primarily limited to the northern Gulf of California. Further
Global mean surface temperature (GMST) Estimated details on how NAmerM is defined and used throughout the Report
global average of near-surface air temperatures over land and sea are provided in Annex V.
ice, and sea surface temperature (SST) over ice-free ocean regions,
with changes normally expressed as departures from a value over South American monsoon (SAmerM)
a specified reference period. See also Global mean surface air The South American monsoon (SAmerM) is a regional circulation
temperature (GSAT). characterized by inflow of low-level winds from the Atlantic to South
America, including Brazil, Peru, Bolivia and northern Argentina,
Global monsoon The global monsoon (GM) is a global-scale associated with the development of surface pressure gradients (and
solstitial mode that dominates the annual variation of tropical and intense precipitation) during austral summer (December–January–
sub-tropical precipitation and circulation. The GM domain is defined February). During September–October–November, areas of intense
as the area where the annual range of precipitation (local summer convection migrate from northwestern South America to the south.
minus winter mean precipitation rate) is greater than 2.5 mm day–1, Associated with this regime, an upper-tropospheric anticyclone
following on from the definition as in Kitoh et al. (2013). Further details (a.k.a. the Bolivian High) forms over the Altiplano region during the
on how the GM is defined, used and related to regional monsoons monsoon onset. The SAmerM then retreats during March–April–May
throughout the Report are provided by Annex V in the AR6 WGI report. with a northeastward migration of the convection. Further details
Australian and Maritime Continent monsoon (AusMCM) on how SAmerM is defined and used throughout the Report are
The Australian–Maritime Continent monsoon (AusMCM) occurs provided in Annex V.
during December–January–February, with the large-scale shift of the South and South East Asian monsoon (SAsiaM)
Inter-tropical Convergence Zone into the Southern Hemisphere and The South and South East Asian monsoon (SAsiaM) is characterized
covering northern Australia and the Maritime Continent up to 10°N. by pronounced seasonal reversals of wind and precipitation. The
The AusMCM is characterized by the seasonal reversal of prevailing AVII
SAsiaM region extends across vast geographical areas and several
easterly winds to westerly winds and the onset of periods of active countries, including India, Bangladesh, Nepal, Myanmar, Sri Lanka,
235
Annex VII Glossary
Pakistan, Thailand, Laos, Cambodia, Vietnam and the Philippines. The equal, the net amount emitted to space is normally less than would
SAsiaM starts in late May/early June and progresses towards the have been emitted in the absence of these absorbers because of
northeast, ending in late September/early October. During the core the decline of temperature with altitude in the troposphere and the
monsoon season, maxima of SAsiaM precipitation are located over consequent weakening of emission. An increase in the concentration
the west coast, north-east and central north India, Myanmar and of GHGs increases the magnitude of this effect; the difference is
Bangladesh, whereas minima are located over north-west and south- sometimes called the enhanced greenhouse effect. The change in
eastern India, western Pakistan, and south-eastern and northern Sri a GHG concentration because of anthropogenic emissions contributes
Lanka. Further details on how SAsiaM is defined and used throughout to an instantaneous radiative forcing. Earth’s surface temperature
the Report are provided in Annex V. and troposphere warm in response to this forcing, gradually restoring
the radiative balance at the top of the atmosphere.
West African monsoon (WAfriM)
The West African monsoon (WAfriM) is a seasonal reversal in wind Greenhouse gas emission metric A simplified relationship
and precipitation whose domain includes Benin, Burkina-Faso, used to quantify the effect of emitting a unit mass of a given
northern Cameroon, Cape Verde, northern Central African Republic, greenhouse gas on a specified key measure of climate change.
Chad, Gambia, Ghana, Guinea, Guinea Bissau, Ivory Coast, Liberia, A relative GHG emission metric expresses the effect from one gas
Mali, Mauritania, Niger, Nigeria, Senegal, Sierra Leone and Togo. The relative to the effect of emitting a unit mass of a reference GHG on
WAfriM is characterized by the northward progression from May to the same measure of climate change. There are multiple emission
September of moist low-level south-westerlies from the Gulf of Guinea. metrics, and the most appropriate metric depends on the application.
In May and June, rainfall essentially remains along the Guinean coast GHG emission metrics may differ with respect to (i) the key measure
with a maximum occurring near 5°N, followed by a sudden decrease of climate change they consider, (ii) whether they consider climate
of rainfall, marking the ‘short dry season‘ in the Guinean coast and outcomes for a specified point in time or integrated over a specified
the monsoon onset in the Sahel. Then rainfall continues to progress time horizon, (iii) the time horizon over which the metric is applied,
northward up to about 18–20°N, with a maximum near 12°N in late (iv) whether they apply to a single emission pulse, emissions sustained
August/September, until it retreats starting from October towards over a period of time, or a combination of both, and (v) whether
the Guinean coast for a second maximum. Further details on how they consider the climate effect from an emission compared to the
WAfriM is defined and used throughout the Report are provided in absence of that emission or compared to a reference emissions level
Annex V. or climate state.
Global surface temperature See Global mean surface Notes: Most relative GHG emission metrics (such as the global
temperature (GMST) and Global mean surface air temperature warming potential (GWP), global temperature change potential
(GSAT). See also Global warming. (GTP), global damage potential, and GWP*) use carbon dioxide (CO2)
as the reference gas. Emissions of non-CO2 gases, when expressed
Global warming Global warming refers to the increase in global
using such metrics, are often referred to as ‘carbon dioxide equivalent’
surface temperature relative to a baseline reference period, averaging
emissions. A metric that establishes equivalence regarding one key
over a period sufficient to remove interannual variations (e.g., 20
measure of the climate system response to emissions does not imply
or 30 years). A common choice for the baseline is 1850–1900 (the
equivalence regarding other key measures. The choice of a metric,
earliest period of reliable observations with sufficient geographic
including its time horizon, should reflect the policy objectives for
coverage), with more modern baselines used depending upon the
which the metric is applied.
application. See also Climate change and Climate variability.
Greenhouse gas neutrality Condition in which metric-
Global warming potential (GWP) An index measuring the
weighted anthropogenic greenhouse gas (GHG) emissions associated
radiative forcing following an emission of a unit mass of a given
with a subject are balanced by metric-weighted anthropogenic
substance, accumulated over a chosen time horizon, relative to that
GHG removals. The subject can be an entity such as a country, an
of the reference substance, carbon dioxide (CO2). The GWP thus
organization, a district or a commodity, or an activity such as a
represents the combined effect of the differing times these substances
service or an event. GHG neutrality is often assessed over the life
remain in the atmosphere and their effectiveness in causing radiative
cycle, including indirect (‘scope 3’) emissions, but can also be limited
forcing. See also Lifetime and Greenhouse gas emission metric.
to the emissions and removals, over a specified period, for which the
Gravitational, rotational and deformational (GRD) subject has direct control, as determined by the relevant scheme.
effects See Sea level change (sea level rise/sea level fall). The quantification of GHG emissions and removals depends on the
GHG emission metric chosen to compare emissions and removals of
Gravity Recovery and Climate Experiment (GRACE) A pair
different gases, as well as the time horizon chosen for that metric.
of satellites that measured the Earth’s gravity field anomalies from
2002 to 2017. These fields have been used, among other things, to Note 1: GHG neutrality and net zero GHG emissions are overlapping
study mass changes of the polar ice sheets and glaciers. concepts. The concepts can be applied at global or sub-global scales
(e.g., regional, national and sub-national). At a global scale, the terms
Greenhouse effect The infrared radiative effect of all infrared-
greenhouse gas neutrality and net zero greenhouse gas emissions are
absorbing constituents in the atmosphere. Greenhouse gases (GHGs),
equivalent. At sub-global scales, net zero greenhouse gas emissions
clouds, and some aerosols absorb terrestrial radiation emitted by the
is generally applied to emissions and removals under direct control
AVII Earth’s surface and elsewhere in the atmosphere. These substances
or territorial responsibility of the reporting entity, while greenhouse
emit infrared radiation in all directions, but, everything else being
gas neutrality generally includes emissions and removals within and
236
Glossary Annex VII
beyond the direct control or territorial responsibility of the reporting Halocarbons A collective term for the group of partially
entity. Accounting rules specified by GHG programmes or schemes halogenated organic species, which includes the chlorofluorocarbons
can have a significant influence on the quantification of relevant (CFCs), hydrochlorofluorocarbons (HCFCs), hydrofluorocarbons
emissions and removals. (HFCs), halons, methyl chloride and methyl bromide. Many of the
halocarbons have large global warming potentials. The chlorine and
Note 2. Under the Paris Rulebook (Decision 18/CMA.1, annex,
bromine-containing halocarbons are also involved in the depletion of
paragraph 37), parties have agreed to use GWP100 values from the
the ozone layer.
IPCC AR5 or GWP100 values from a subsequent IPCC Assessment
Report to report aggregate emissions and removals of GHGs. In Halocline A layer in the oceanic water column in which salinity
addition, parties may use other metrics to report supplemental changes rapidly with depth. Generally, saltier water is denser and
information on aggregate emissions and removals of GHGs. lies below less salty water. In some high-latitude oceans the surface
waters may be colder than the deep waters, and the halocline is
Note 3: In some cases, achieving greenhouse gas neutrality may rely
responsible for maintaining water column stability and isolating the
on the supplementary use of offsets to balance emissions that remain
surface waters from the deep waters.
after actions by the reporting entity are taken into account.
Halosteric See Sea level change (sea level rise/sea level fall).
See also Carbon neutrality, Greenhouse gas emission metric and Net
zero greenhouse gas emissions. Halosteric sea level change See Sea level change (sea level
rise/sea level fall).
Greenhouse gases (GHGs) Gaseous constituents of the
atmosphere, both natural and anthropogenic, that absorb and emit Hazard The potential occurrence of a natural or human-induced
radiation at specific wavelengths within the spectrum of radiation physical event or trend that may cause loss of life, injury, or other
emitted by the Earth’s surface, by the atmosphere itself, and by health impacts, as well as damage and loss to property, infrastructure,
clouds. This property causes the greenhouse effect. Water vapour livelihoods, service provision, ecosystems and environmental
(H2O), carbon dioxide (CO2), nitrous oxide (N2O), methane (CH4 ) and resources. See also Impacts (consequences, outcomes) and Risk.
ozone (O3) are the primary GHGs in the Earth’s atmosphere. Human-
Heat index A measure of how hot the air feels to the human body.
made GHGs include sulphur hexafluoride (SF6 ), hydrofluorocarbons
The index is mainly based on surface air temperature and relative
(HFCs), chlorofluorocarbons (CFCs) and perfluorocarbons (PFCs);
humidity; thus it reflects the combined effect of high temperature and
several of these are also O3-depleting (and are regulated under the
humidity on human physiology and provides a relative indication of
Montreal Protocol). See also Well-mixed greenhouse gas.
potential health risks.
Greenland Ice Sheet (GrIS) See Ice sheet.
Heat stress A range of conditions in, for example, terrestrial
Gross Primary Production (GPP) See Primary production. or aquatic organisms when the body absorbs excess heat during
overexposure to high air or water temperatures or thermal radiation.
Ground-level ozone Atmospheric ozone (O3) is formed naturally
In aquatic water-breathing animals, hypoxia and acidification can
or from human-emitted precursors near Earth’s surface, thus affecting
exacerbate vulnerability to heat. Heat stress in mammals (including
human health, agriculture and ecosystems. Ozone is a greenhouse
humans) and birds, both in air, is exacerbated by a detrimental
gas (GHG), but ground-level ozone, unlike stratospheric ozone,
combination of ambient heat, high humidity and low wind speeds,
also directly affects organisms at the surface. Ground-level ozone
causing regulation of body temperature to fail.
is sometimes referred to as tropospheric ozone, although much of
the troposphere is well above the surface and thus does not directly Heatwave A period of abnormally hot weather, often defined with
expose organisms at the surface. reference to a relative temperature threshold, lasting from two days
to months. Heatwaves and warm spells have various and, in some
Grounding line The junction between a glacier or ice sheet and
cases, overlapping definitions. See also Marine heatwave, Blocking,
an ice shelf; the place where ice starts to float. This junction normally
Heat index and Heat stress.
occurs over a zone, rather than at a line.
Heavy precipitation event See Extreme/heavy precipitation
Gyre Basin-scale ocean horizontal circulation pattern with slow
event.
flow circulating around the ocean basin, closed by a strong and
narrow (100 to 200 km wide) boundary current on the western side. Heinrich event Distinct layers of coarse-grained sediments
The subtropical gyres in each ocean are associated with high pressure comprised of ice-rafted debris identified across marine sediment cores
in the centre of the gyres; the subpolar gyres are associated with low in the North Atlantic. These sedimentary layers are closely associated
pressure. with millennial-scale cooling events in the North Atlantic and a
distinct pattern of global temperature and hydrological changes that
Hadley cell See Hadley circulation.
are largely consistent with evidence for a slowdown, or even near-
Hadley circulation A direct, thermally driven overturning cell in collapse, of the Atlantic Meridional Overturning Circulation (AMOC)
the atmosphere consisting of poleward flow in the upper troposphere, during these times.
subsiding air into the subtropical anticyclones, return flow as part of
Heterotrophic respiration The conversion of organic matter to
the trade winds near the surface, and with rising air near the equator
carbon dioxide (CO2) by organisms other than autotrophs.
in the so-called Inter-tropical Convergence Zone. AVII
237
Annex VII Glossary
Holocene The current interglacial geological epoch, the second recharges groundwater, discharges into streams, and ultimately, flows
of two epochs within the Quaternary Period, the preceding being into the oceans as rivers, polar glaciers and ice sheets, from which it
the Pleistocene. The International Commission on Stratigraphy (ICS) will eventually evaporate again. The various systems involved in the
defines the start of the Holocene Epoch at 11,700 years before 2000 hydrological cycle are usually referred to as hydrological systems.
(Walker et al., 2019) spanning the interval from 11,700 yr to the
Hydrological drought See Drought.
present day. Together with the subadjacent Pleistocene, it comprises
the Quaternary System/Period. The Holocene record contains diverse Hydrological sensitivity (η) The linear change in global
geomorphological, biological, climatological and archaeological mean precipitation per degree Celsius of global mean surface air
evidence, within sequences that are often continuous and extremely temperature (GSAT) change once precipitation changes related to
well-preserved at decadal, annual and even seasonal resolution. fast atmospheric and land surface adjustments to radiative forcings
As a consequence, the Holocene is perhaps the most intensively have occurred. Units are % per °C although it can also be calculated
studied series/epoch within the entire Geological Time Scale. Yet as W m–2 per °C. See also Apparent hydrological sensitivity (ηa ).
until recently little attention had been paid to a formal subdivision of
Hydrosphere The component of the climate system comprising
the Holocene. Here we describe an initiative by the Subcommission
liquid surface and subterranean water, such as in oceans, seas, rivers,
on Quaternary Stratigraphy (SQS. It encompasses the mid-Holocene
freshwater lakes, underground water, wetlands, etc.
(MH), the 1000-year-long interval centred at 6000 years before 1950;
a period of long-standing focus for climate modelling, with enhanced Hypoxic Conditions of low dissolved oxygen in shallow water
seasonality in the Northern Hemisphere and decreased seasonality in ocean and freshwater environments. There is no universal threshold
the Southern Hemisphere. The early part of the Holocene is marked by for hypoxia. A value around 60 μmol kg–1 has commonly been used
the late stages of deglaciation of Pleistocene land ice, sea level rise, for some estuarine systems, although this does not necessarily
and the occurrence of warm phases that affected different regions at directly translate into biological impacts. Anoxic conditions occur
different times, often referred to as the ‘Holocene Thermal Maximum’. where there is no oxygen present at all. See also Eutrophication.
In addition, the epoch includes the post-glacial interval, which began
approximately 7000 years ago when the fundamental features of the Hypsometry The distribution of land or ice surface as a function
modern climate system were essentially in place, as the influence of of altitude.
remnant Pleistocene ice sheets waned. See also Anthropocene. Ice age An informal term for a geological period characterized
Holocene Thermal Maximum (HTM) See Holocene. by a long-term reduction in the temperature of the Earth’s climate,
resulting in the presence or expansion of ice sheets and glaciers.
Human influence on the climate system Human-driven Among the Earth’s ice ages is the current Quaternary Period,
activities that lead to changes in the climate system due to characterized by alternating glacial and interglacial intervals. See
perturbations of the Earth’s energy budget (also called anthropogenic also Deglacial or deglaciation or glacial termination and Glacial–
forcing). Human influence results from emissions of greenhouse interglacial cycles.
gases, aerosols, ozone-depleting substances (ODSs), and land-use
change. See also Anthropogenic, Anthropogenic emissions and Ice core A cylinder of ice drilled out of a glacier or ice sheet
Anthropogenic removals. to determine the physical properties of the ice body and to gain
information on past changes in climate and composition of the
Human system Any system in which human organizations atmosphere that are preserved in the ice or in air trapped in the ice.
and institutions play a major role. Often, but not always, the term
is synonymous with society or social system. Systems such as Ice sheet An ice body originating on land that covers an area of
agricultural systems, urban systems, political systems, technological continental size, generally defined as covering >50,000 km2, and
systems and economic systems are all human systems in the sense that has formed over thousands of years through accumulation and
applied in this Report. compaction of snow. An ice sheet flows outward from a high central
ice plateau with a small average surface slope. The margins usually
Hurricane See Tropical cyclone. slope more steeply, and most ice is discharged through fast-flowing
ice streams or outlet glaciers, often into the sea or into ice shelves
Hydroclimate Part of the climate pertaining to the hydrology of
floating on the sea. There are only two ice sheets in the modern
a region.
world, one on Greenland and one on Antarctica. The latter is divided
Hydrofluorocarbons (HFCs) A type of greenhouse gas (GHG), into the East Antarctic Ice Sheet (EAIS), the West Antarctic Ice Sheet
HFCs are organic compounds that contain fluorine, carbon and (WAIS) and the Antarctic Peninsula Ice Sheet. During glacial periods,
hydrogen atoms and they are produced commercially as a substitute there were other ice sheets.
for chlorofluorocarbons (CFCs). They are mainly used in refrigeration
Ice shelf A floating slab of ice originating from land of considerable
and semiconductor manufacturing.
thickness extending from the coast (usually of great horizontal extent
Hydrological cycle The cycle in which water evaporates from the with a very gently sloping surface), resulting from the flow of ice
ocean and the land surface, is carried over the Earth in atmospheric sheets, initially formed by the accumulation of snow, and often filling
circulation as water vapour, condenses to form clouds, precipitates embayments in the coastline of an ice sheet. Nearly all ice shelves are
over the ocean and land as rain or snow, which on land can be in Antarctica, where most of the ice discharged into the ocean flows
AVII intercepted by trees and vegetation, potentially accumulating as via ice shelves.
snow or ice, provides runoff on the land surface, infiltrates into soils,
238
Glossary Annex VII
Ice stream A stream of ice with strongly enhanced flow that is during the second half of the 18th century and spreading to Europe
part of an ice sheet. It is often separated from surrounding ice by and later to other countries including the United States. The invention
strongly sheared, crevassed margins. of the steam engine was an important trigger of this development.
The industrial revolution marks the beginning of a strong increase
Ice–albedo feedback A climate feedback involving changes
in the use of fossil fuels, initially coal, and hence emission of carbon
in the Earth’s surface albedo. Snow and ice have an albedo much
dioxide (CO2).
higher (up to ~0.8) than the average planetary albedo (~0.3). With
increasing temperatures, it is anticipated that snow and ice extent Infrared radiation See Terrestrial radiation.
will decrease, the Earth’s overall albedo will decrease and more solar
Initial condition ensemble (ICE) See Climate simulation
radiation will be absorbed, warming the Earth further.
ensemble.
Iceberg Large piece of freshwater ice broken off from a glacier
Insolation The amount of solar radiation reaching the Earth by
or an ice shelf during calving and floating in open water (at least
latitude and by season measured in W m–2. Usually, insolation refers
5 m height above sea level). Smaller pieces of floating ice known as
to the radiation arriving at the top of the atmosphere. Sometimes it is
‘bergy bits’ (less than 5 m above sea level) or ‘growlers’ (less than
specified as referring to the radiation arriving at the Earth’s surface.
2 m above sea level) can originate from glaciers or ice shelves, or from
See also Orbital forcing and Total solar irradiance (TSI).
the breaking up of a large iceberg. Icebergs can also be classified by
shape, most commonly being either tabular (steep sides and a flat Instantaneous radiative forcing (or effect) due to aerosol–
top) or non-tabular (varying shapes, with domes and spires) (NOAA, cloud interactions (IRFaci) See Aerosol–cloud interaction.
2021). In lakes, icebergs can originate by breaking off shelf ice, which
Instantaneous radiative forcing (or effect) due to aerosol–
forms through freezing of a lake surface.
radiation interactions (IRFari) See Aerosol–radiation
Impacts The consequences of realized risks on natural and human interaction.
systems, where risks result from the interactions of climate-related
Integrated assessment model (IAM) Models that integrate
hazards (including extreme weather/climate events), exposure, and
knowledge from two or more domains into a single framework. They
vulnerability. Impacts generally refer to effects on lives, livelihoods,
are one of the main tools for undertaking integrated assessments.
health and well-being, ecosystems and species, economic, social
One class of IAM used with respect to climate change mitigation may
and cultural assets, services (including ecosystem services), and
include representations of: multiple sectors of the economy, such as
infrastructure. Impacts may be referred to as consequences or
energy, land use and land-use change; interactions between sectors;
outcomes and can be adverse or beneficial. See also Adaptation,
the economy as a whole; associated greenhouse gas (GHG) emissions
Exposure, Hazard, Vulnerability and Risk.
and sinks; and reduced representations of the climate system. This
Incoming solar radiation See Insolation. class of model is used to assess linkages between economic, social
and technological development and the evolution of the climate
Indian Ocean basin (IOB) mode A mode of interannual
system. Another class of IAM additionally includes representations of
variability characterized by a temporal alternation of basin-wide
the costs associated with climate change impacts, but includes less
warming and cooling of the Indian Ocean sea surface. It mostly
detailed representations of economic systems. These can be used to
develops in response to El Niño–Southern Oscillation (ENSO), but
assess impacts and mitigation in a cost–benefit framework and have
often persists after ENSO’s equatorial eastern Pacific signal has
been used to estimate the social cost of carbon.
dissipated. The IOB affects atmospheric circulation, temperature, and
precipitation in South, South East, and East Asia as well as Africa, and Inter-decadal Pacific Oscillation (IPO) See Pacific Decadal
modulates tropical cyclone activity in the north-western Pacific. See Variability (PDV).
Section AIV.2.4 in Annex IV of the AR6 WGI report. See also Modes of
Inter-tropical Convergence Zone (ITCZ) The Inter-tropical
climate variability and Indian Ocean Dipole (IOD).
Convergence Zone is an equatorial zonal belt of low pressure, strong
Indian Ocean Dipole (IOD) A mode of interannual variability that convection and heavy precipitation near the equator where the
features an east–west dipole of sea surface temperature anomalies north-east trade winds meet the south-east trade winds. This band
in the tropical Indian Ocean. Its positive phase shows concurrent sea moves seasonally. See also South Pacific Convergence Zone (SPCZ).
surface cooling off Sumatra and Java and warming off Somalia in the
Interglacial or interglaciation A globally warm period lasting
west, combined with anomalous surface easterlies along the equator,
thousands of years between glacial periods within an ice age.
while the opposite anomalies are seen in the negative phase. The IOD
Generally coincides with odd-numbered marine isotope stages (MIS)
typically develops in boreal summer and matures in boreal autumn
when mean sea level was close to present. The Last Interglacial (LIG)
and controls part of the rainfall interannual variability in Australia,
occurred between about 129 and 116 ka (thousand years) before
South Eastern Asia and Eastern Africa. See Section AIV.2.4 in Annex
present (defined as 1950) although the warm period started in some
IV of the AR6 WGI report. See also Indian Ocean Basin (IOB) mode.
areas a few thousand years earlier. In terms of MIS, interglaciations
Indirect aerosol effect See Aerosol–cloud interaction. are defined as the interval between the midpoint of the preceding
termination and the onset of the next glaciation. The LIG coincides
Indirect land-use change (iLUC) See Land-use change (LUC).
with MIS 5e. The present interglaciation, the Holocene, started at
Industrial revolution A period of rapid industrial growth with 11,700 years before 2000 CE, although global mean sea level did not AVII
far-reaching social and economic consequences, beginning in Britain approach its present position until roughly 7000 years ago. See also
239
Annex VII Glossary
Deglacial or deglaciation or glacial termination, Glacial-interglacial Land–cover change Change from one land cover class to
cycles, Glacial or glaciation and Ice age. another, due to change in land use or change in natural conditions
(Pongratz et al., 2018). See also Land-use change (LUC).
Internal climate variability See Internal variability (under
Climate variability). Land surface air temperature (LSAT) The near-surface air
temperature over land, typically measured at 1.25–2 m above the
Interstadial or interstade A brief period of regional climatic
ground using standard meteorological equipment.
warming during a glacial or interglacial interval, often characterized
by transient glacial retreats. Interstadials are generally of short Land use The total of arrangements, activities and inputs applied
duration (hundreds to a few thousand years) compared to glacial or to a parcel of land. The term land use is also used in the sense of
interglacial intervals (lasting many thousands to tens of thousands the social and economic purposes for which land is managed (e.g.,
of years). One example of a regional interstadial event is based grazing, timber extraction, conservation and city dwelling). In
on millennial scale warming recorded by oxygen isotope ratios in national greenhouse gas (GHG) inventories, land use is classified
Greenland ice cores, the so called “Greenland Interstadials” (Johnsen according to the IPCC land-use categories of forest land, cropland,
et al., 1992). See also Stadial or stade. grassland, wetlands, settlements, other lands (see the 2006 IPCC
Guidelines for National GHG Inventories and their 2019 Refinement
Irreversibility A perturbed state of a dynamical system is defined
for details (IPCC, 2006, 2019)).
as irreversible on a given time scale if the recovery from this state due
to natural processes takes substantially longer than the time scale of Land-use change (LUC) The change from one land use category
interest. See also Tipping point. to another. Note that in some scientific literature, land-use change
encompasses changes in land-use categories as well as changes
Isostatic or Isostasy Isostasy refers to the response of the
in land management. See also Afforestation, Deforestation and
Earth to changes in surface load. It includes the deformational and
Reforestation.
gravitational response. This response is elastic on short time scales,
as in the Earth–ocean response to recent changes in mountain Indirect land-use change (iLUC) Land-use change outside the
glaciation, or viscoelastic on longer time scales, as in the response to area of focus that occurs as a consequence of change in use or
the last deglaciation following the Last Glacial Maximum. management of land within the area of focus, such as through market
or policy drivers. For example, if agricultural land is diverted to biofuel
Isotopes Atoms of the same chemical element that have the
production, forest clearance may occur elsewhere to replace the
same the number of protons but differ in the number of neutrons.
former agricultural production. See Land-use change (LUC).
Some proton–neutron configurations are stable (stable isotopes),
others are unstable undergoing spontaneous radioactive decay Land water storage (LWS) Land water storage (LWS) includes
(radioisotopes). Most elements have more than one stable isotope. all surface water, soil moisture, groundwater storage and snow, but
Isotopes can be used to trace transport processes or to study excludes water stored in glaciers and ice sheets. Changes in LWS can
processes that change the isotopic ratio. Radioisotopes provide, in be caused either by direct human intervention in the water cycle (e.g.,
addition, time information that can be used for radiometric dating. storage of water in reservoirs by building dams in rivers, groundwater
See also 13C and 14C. extraction from groundwater reservoirs for consumption and
irrigation, or deforestation) or by climate variations (e.g., changes in
Key climate indicators See Climate indicator.
the amount of water in endorheic lakes and wetlands, the canopy,
Kriging Kriging is a method of interpolation (normally spatial the soil, the permafrost and the snowpack). Land water storage
interpolation when used with atmospheric or oceanographic data) changes caused by climate variations may also be indirectly affected
in which the interpolated values are estimated using a Gaussian by anthropogenic influences. See also Sea level change (sea level rise/
process governed by prior covariances. sea level fall).
La Niña See El Niño–Southern Oscillation (ENSO). Lapse rate The rate of change of an atmospheric variable, usually
temperature, with height. The lapse rate is considered positive when
Land The terrestrial portion of the biosphere that comprises the
the variable decreases with height.
natural resources (soil, near-surface air, vegetation and other biota,
and water), the ecological processes, topography, and human Large-scale The climate system involves process interactions
settlements and infrastructure that operate within that system from the micro- to the global-scale. Any threshold for defining ‘large-
(UNCCD, 1994; FAO, 2007). scale’ is arbitrary. Understanding of large-scale climate variability
and change requires knowledge of both the response to external
Land cover The biophysical coverage of land (e.g., bare soil, rocks,
forcings and the role of internal variability. Many external forcings
forests, buildings and roads or lakes). Land cover is often categorized
have substantial hemispheric or continental scale variations. Modes
in broad land-cover classes (e.g., deciduous forest, coniferous forest,
of climate variability are driven by ocean-basin-scale processes. Thus
mixed forest, grassland, bare ground). [Note: In some literature, land
we define large-scale to include ocean-basin and continental scales
cover and land use are used interchangeably, but the two represent
as well as hemispheric and global scales.
distinct classification systems. For example, the land cover class
woodland can be under various land uses such as livestock grazing, Last deglacial transition See Deglacial or deglaciation or
recreation, conservation, or wood harvest.] glacial termination and Younger Dryas.
AVII
240
Glossary Annex VII
Last Glacial Maximum (LGM) See Glacial or glaciation. Turnover time (T) (also called global atmospheric lifetime) is the
ratio of the mass M of a reservoir (e.g., a gaseous compound in
Last Interglacial (LIG) See Interglacial or interglaciation.
the atmosphere) and the total rate of removal S from the reservoir:
Last millennium The interval of the Common Era (CE) between T = M/S. For each removal process, separate turnover times can
1001 and 2000 CE. Encompasses the Little Ice Age, a roughly defined be defined. In soil carbon biology, this is referred to as mean
period characterized by multiple expansions of mountain glaciers residence time.
worldwide, the timing of which differs among regions but generally
Light-absorbing particles Light-absorbing particles (LAP), for
occurred between 1400 CE and 1900 CE. The last millennium also
example, black carbon (BC), brown carbon and dust, are particles
mostly encompasses the Medieval Warm Period (also called the
that absorb solar radiation and convert it into internal energy, thus
Medieval Climate Anomaly), a roughly defined period of relatively
raising the particle’s temperature and emitting thermal-infrared
warm conditions or other climate excursions such as extensive
radiation that is selectively absorbed by the surrounding medium.
drought, the timing and magnitude of which differ among regions,
LAP affect the energy balance of the atmosphere and clouds, and
but generally occurred between 900 and 1400 CE. Transient climate
when deposited on snow and ice, they reduce snow/ice albedo,
model experiments by the Paleoclimate Modelling Intercomparison
increasing heating and accelerating melting. These particles have a
Project (PMIP) for the last millennium extend from 850–1849 CE.
warming effect on climate.
Latent heat flux The turbulent flux of heat from the Earth’s
Likelihood The chance of a specific outcome occurring, where
surface to the atmosphere that is associated with evaporation or
this might be estimated probabilistically. Likelihood is expressed in
condensation of water vapour at the surface; a component of the
this report using a standard terminology (Mastrandrea et al., 2010).
surface energy budget. See also Sensible heat flux.
See also Agreement, Confidence, Evidence and Uncertainty.
Lifetime Lifetime is a general term used for various time scales
Lithosphere The upper layer of the solid Earth, both continental
characterizing the rate of processes affecting the concentration of
and oceanic, which comprises all crustal rocks and the cold, mainly
trace gases. The following lifetimes may be distinguished:
elastic part of the uppermost mantle. Volcanic activity, although part
Response time or adjustment time (Ta ) of the lithosphere, is not considered as part of the climate system, but
Response time or adjustment time (Ta) is the time scale characterizing acts as an external forcing factor.
the decay of an instantaneous pulse input into the reservoir. The term
Livelihood The resources used and the activities undertaken in
adjustment time is also used to characterize the adjustment of the
order for people to live. Livelihoods are usually determined by the
mass of a reservoir following a step change in the source strength.
entitlements and assets to which people have access. Such assets can
Half-life or decay constant is used to quantify a first-order exponential
be categorized as human, social, natural, physical or financial.
decay process. See Response time or adjustment time for a different
definition pertinent to climate variations. Local sea level change See Sea level change (sea level rise/sea
level fall).
The term lifetime is sometimes used, for simplicity, as a surrogate for
adjustment time. Long-lived greenhouse gases (LLGHGs) A set of well-mixed
greenhouse gases with long atmospheric lifetimes. This set of
In simple cases, where the global removal of the compound is directly
compounds includes carbon dioxide (CO2) and nitrous oxide (N2O),
proportional to the total mass of the reservoir, the adjustment time
together with some halogenated compounds. They have a warming
equals the turnover time: T = Ta. An example is CFC-11, which is
effect on climate. These compounds accumulate in the atmosphere at
removed from the atmosphere only by photochemical processes in
decadal to centennial time scales, and their effect on climate hence
the stratosphere. In more complicated cases, where several reservoirs
persists for decades to centuries after their emission. On time scales
are involved or where the removal is not proportional to the total
of decades to a century, already emitted emissions of long-lived
mass, the equality T = Ta no longer holds.
climate forcers can only be abated by greenhouse gas removal.
Carbon dioxide (CO2) is an extreme example. Its turnover time
Longwave radiation See Terrestrial radiation.
is only about 4 years because of the rapid exchange between the
atmosphere and the ocean and terrestrial biota. However, a large part Low-likelihood, high impact outcomes Outcomes/events
of that CO2 is returned to the atmosphere within a few years. The whose probability of occurrence is low or not well known (as in the
adjustment time of CO2 in the atmosphere is determined from the context of deep uncertainty) but whose potential impacts on society
rates of removal of carbon by a range of processes with time scales and ecosystems could be high. To better inform risk assessment and
from months to hundreds of thousands of years. As a result, 15 to decision-making, such low-likelihood outcomes are considered if
40% of an emitted CO2 pulse will remain in the atmosphere longer they are associated with very large consequences and may therefore
than 1,000 years, 10 to 25% will remain about ten thousand years, constitute material risks, even though those consequences do not
and the rest will be removed over several hundred thousand years. necessarily represent the most likely outcome.
In the case of methane (CH4 ), the adjustment time is different from Madden–Julian Oscillation (MJO) The largest mode of
the turnover time because the removal is mainly through a chemical tropical atmospheric intra-seasonal variability with typical periods
reaction with the hydroxyl radical (OH), the concentration of which ranging from 20 to 90 days. The MJO corresponds to planetary-
itself depends on the CH4 concentration. Therefore, the CH4 removal scale disturbances of pressure, wind and deep convection moving AVII
rate S is not proportional to its total mass M. predominantly eastward along the equator. As it progresses, the MJO
241
Annex VII Glossary
is associated with the temporal alternation of large-scale enhanced Accumulation (of glaciers, ice sheets, or snow cover)
and suppressed rainfall, with maximum loading over the Indian All processes that add to the mass of a glacier, an ice sheet, or snow
and western Pacific oceans, although influences of the MJO can cover. The main process of accumulation is snowfall. Accumulation
be tracked over the Atlantic/Africa in dynamical fields. See Section also includes deposition of hoar, freezing rain, other types of solid
AIV.2.8 in Annex IV of the AR6 WGI report. precipitation, gain of wind-blown snow, avalanching, and basal
accumulation (often beneath floating ice).
Maladaptive actions (Maladaptation) Actions that may lead
to increased risk of adverse climate-related outcomes, including via Discharge (of ice)
increased greenhouse gas (GHG) emissions, increased vulnerability Rate of the flow of ice through a vertical section of a glacier
to climate change, or diminished welfare, now or in the future. perpendicular to the direction of the flow of ice. Often used to refer
Maladaptation is usually an unintended consequence. to the loss of mass at marine-terminating glacier fronts (mostly
calving of icebergs and submarine melt), or to mass flowing across
Marine cloud brightening (MCB) See Solar radiation
the grounding line of a floating ice shelf.
modification (SRM).
Mean sea level The surface level of the ocean at a particular
Marine heatwave A period during which water temperature
point averaged over an extended period of time such as a month
is abnormally warm for the time of the year relative to historical
or year. Mean sea level is often used as a national datum to which
temperatures, with that extreme warmth persisting for days to
heights on land are referred.
months. The phenomenon can manifest in any place in the ocean and
at scales of up to thousands of kilometres. See also Heatwave. Megacity Urban agglomerations with 10 million inhabitants
or more.
Marine ice cliff instability (MICI) A hypothetical mechanism
of an ice cliff failure. In case a marine-terminated ice sheet loses its Meltwater Pulse 1A (MWP-1A) A particular interval of rapid
buttressing ice shelf, an ice cliff can be exposed. If the exposed ice global sea level rise between about 14,700 and 14,300 years
cliff is tall enough (about 800 m of the total height, or about 100 ago, associated with the end of the last ice age and attributed to
m of the above-water part), the stresses at the cliff face exceed the freshwater flux to the ocean from accelerated melting of ice sheets
strength of the ice, and the cliff fails structurally in repeated calving and glaciers. First defined based on oxygen isotope data (Duplessy
events. See also Marine ice sheet instability (MISI). et al., 1981), and later shown to be reflected by high rates of sea
level rise (Fairbanks, 1989). See also Deglacial or deglaciation or
Marine ice sheet instability (MISI) A mechanism of
glacial termination.
irreversible (on the decadal to centennial time scale) retreat of a
grounding line for the marine-terminating glaciers, in case the glacier Meridional overturning circulation (MOC) Meridional
bed slopes towards the ice sheet interior. See also Marine ice cliff (north–south) overturning circulation in the ocean quantified by
instability (MICI). zonal (east–west) sums of mass transports in depth or density
layers. In the North Atlantic, away from the subpolar regions, the
Marine isotope stage (MIS) Geological periods of alternating
MOC (which is in principle an observable quantity) is often identified
glacial and interglacial conditions, each typically lasting tens of
with the thermohaline circulation (THC), which is a conceptual and
thousands of years as inferred from the oxygen isotope composition
incomplete interpretation. The MOC is also driven by wind, and can
of microfossils from deep sea sediment cores. MIS numbers
also include shallower overturning cells such as occur in the upper
increase back in time from the present, which is MIS 1. Even-
ocean in the tropics and subtropics, in which warm (light) waters
number MISs coincide with glacial periods, and odd-numbered MISs
moving poleward are transformed to slightly denser waters and
are interglacials.
subducted equatorward at deeper levels.
Marine-based ice sheet An ice sheet containing a substantial
Atlantic Meridional Overturning Circulation (AMOC)
region that rests on a bed lying below sea level and whose perimeter
The main current system in the South and North Atlantic Oceans.
is in contact with the ocean. The best known example is the West
AMOC transports warm upper-ocean water northwards and cold,
Antarctic Ice Sheet.
deep water southwards, as part of the global ocean circulation system.
Mass balance/budget (of glaciers or ice sheets) Difference Changes in the strength of AMOC can affect other components of the
between the mass input (accumulation) and the mass loss (ablation) climate system.
of an ice body (e.g., a glacier or ice sheet) over a stated time period,
Meteorological drought See Drought.
which is often a year or a season. Surface mass balance refers to the
difference between surface accumulation and surface ablation. Methane (CH4) The greenhouse gas methane is the major
component of natural gas and associated with all hydrocarbon fuels.
Ablation (of glaciers, ice sheets, or snow cover)
Significant anthropogenic emissions also occur as a result of animal
All processes that reduce the mass of a glacier, ice sheet, or snow
husbandry and paddy rice production. Methane is also produced
cover. The main processes are melting, and for glaciers also calving
naturally where organic matter decays under anaerobic conditions,
(or, when the glacier nourishes an ice shelf, discharge of ice across
such as in wetlands. Under future global warming, there is potential
the grounding line), but other processes such as sublimation and loss
for increased methane emissions from thawing permafrost, wetlands
of wind-blown snow can also contribute to ablation. Ablation also
and sub-sea gas hydrates. See also Short-lived climate forcers (SLCFs).
AVII refers to the mass lost by any of these processes.
Microclimate Local climate at or near the Earth’s surface.
242
Glossary Annex VII
Microwave sounding unit (MSU) A microwave sounder on progresses, thereby limiting the time period over which the forecast
U.S. National Oceanic and Atmospheric Administration (NOAA) polar will be useful.
orbiter satellites that estimates the temperature of thick layers of the
Model spread The range or spread in results from climate models,
atmosphere by measuring the thermal emission of oxygen molecules
such as those assembled for Coupled Model Intercomparison Project
from a complex of emission lines near 60 GHz. A series of nine MSUs
Phase 6 (CMIP6). Does not necessarily provide an exhaustive and
began making this kind of measurement in late 1978. Beginning in
formal estimate of the uncertainty in feedbacks, forcing or projections
mid-1998, a follow-on series of instruments, the Advanced Microwave
even when expressed numerically, for example, by computing
Sounding Units (AMSUs), began operation.
a standard deviation of the models’ responses. In order to quantify
Mid-Holocene (MH) See Holocene. uncertainty, information from observations, physical constraints and
expert judgement must be combined, using a statistical framework.
Mid-Pliocene Warm Period (MPWP) See Pliocene.
Modes of climate variability Recurrent space-time structures
Mineralization/Remineralization The conversion of an
of natural variability of the climate system with intrinsic spatial
element from its organic form to an inorganic form as a result
patterns, seasonality and time scales. Modes can arise through the
of microbial decomposition. In nitrogen mineralization, organic
dynamical characteristics of the atmospheric circulation but also
nitrogen from decaying plant and animal residues (proteins, nucleic
through coupling between the ocean and the atmosphere, with some
acids, amino sugars and urea) is converted to ammonia (NH3) and
interactions with land surfaces and sea ice. Many modes of variability
ammonium (NH4+) by biological activity.
are driven by internal climate processes and are a critical potential
Mitigation (of climate change) A human intervention to source of climate predictability on sub-seasonal to decadal time
reduce emissions or enhance the sinks of greenhouse gases. scales. See Annex IV of the AR6 WGI report. See also Annular modes,
Tropical Atlantic Variability (TAV), Indian Ocean Dipole (IOD), Indian
Mitigation pathways See Pathways.
Ocean Basin (IOB) mode, Pacific Decadal Variability (PDV), Pacific
Mitigation potential The quantity of net greenhouse gas Decadal Oscillation (PDO) (under Pacific Decadal Variability (PDV)),
emission reductions that can be achieved by a given mitigation El Niño–Southern Oscillation (ENSO), North Atlantic Oscillation
option relative to specified emission baselines. [Note: Net greenhouse (NAO), Northern Annular Mode (NAM) (under Annular modes),
gas emission reductions is the sum of reduced emissions and/or Southern Annular Mode (SAM) (under Annular modes), Atlantic
enhanced sinks.] See also Sequestration potential. Meridional Mode (AMM) (under Tropical Atlantic Variability (TAV)),
Atlantic Zonal Mode (AZM) (under Tropical Atlantic Variability (TAV)),
Biogeophysical potential Madden–Julian Oscillation (MJO), Atlantic Multi-decadal Variability
The mitigation potential constrained by biological, geophysical and (AMV) and Inter-decadal Pacific Oscillation (IPO) (under Pacific
geochemical limits and thermodynamics, without taking into account Decadal Variability (PDV)).
technical, social, economic and/or environmental considerations.
Mole fraction or mixing ratio Mole fraction, or mixing ratio, is
Economic potential the ratio of the number of moles of a constituent in a given volume
The portion of the technical potential for which the social benefits to the total number of moles of all constituents in that volume. It is
exceed the social costs, taking into account a social discount rate and usually reported for dry air. Typical values for well-mixed greenhouse
the value of externalities. gases are in the order of μmol mol–1 (parts per million: ppm), nmol
Technical potential mol–1 (parts per billion: ppb), and fmol mol–1 (parts per trillion: ppt).
The mitigation potential constrained by biogeophysical limits as well as Mole fraction differs from volume mixing ratio, often expressed in
availability of technologies and practices. Quantification of technical ppmv, etc., by the corrections for non-ideality of gases. This correction
potentials takes into account primarily technical considerations, is significant relative to measurement precision for many greenhouse
but social, economic and/or environmental considerations are gases (Schwartz and Warneck, 1995).
occasionally also included, if these represent strong barriers for the Monsoon See Global monsoon.
deployment of an option.
Montreal Protocol The Montreal Protocol on Substances that
Mitigation scenario See Scenario. Deplete the Ozone Layer was adopted in Montreal in 1987, and
Mixing ratio See Mole fraction or mixing ratio. subsequently adjusted and amended (including London (1990),
Copenhagen (1992), Vienna (1995), Montreal (1997), Beijing (1999)
Model ensemble See Climate simulation ensemble. See and Kigali (2016)). It controls the consumption and production of
also Ensemble. chlorine- and bromine-containing chemicals that destroy stratospheric
Model initialization A climate prediction typically proceeds ozone (O3 ), such as chlorofluorocarbons (CFCs), methyl chloroform,
by integrating a climate model forward in time from an initial state carbon tetrachloride and many others. Since the Kigali Amendment in
that is intended to reflect the actual state of the climate system. 2016, hydrofluorocarbons (HFCs), which were used as alternatives to
Available observations of the climate system are assimilated into the ozone-depleting substances (ODSs), have been targeted for a phase-
model. Initialization is a complex process that is limited by available down due to their climate effect as greenhouse gases (GHGs).
observations, observational errors and, depending on the procedure Multi-model ensemble (MME) See Climate simulation
used, may be affected by uncertainty in the history of climate forcing. AVII
ensemble. See also Ensemble.
The initial conditions will contain errors that grow as the forecast
243
Annex VII Glossary
Narrative See Storyline. See also Pathways. emissions is generally applied to emissions and removals under
direct control or territorial responsibility of the reporting entity,
Natural systems The dynamic physical, physicochemical and
while GHG neutrality generally includes anthropogenic emissions
biological components of the Earth system that would operate
and anthropogenic removals within and beyond the direct control
independently of human activities.
or territorial responsibility of the reporting entity. Accounting rules
Natural variability See Climate variability. specified by GHG programmes or schemes can have a significant
influence on the quantification of relevant emissions and removals.
Near-surface permafrost See Permafrost.
Note 2: Under the Paris Rulebook (Decision 18/CMA.1, annex,
Negative greenhouse gas emissions Removal of greenhouse
paragraph 37), parties have agreed to use GWP100 values from the
gases (GHGs) from the atmosphere by deliberate human activities,
IPCC AR5 or GWP100 values from a subsequent IPCC Assessment
that is, in addition to the removal that would occur via natural carbon
Report to report aggregate emissions and removals of GHGs. In
cycle or atmospheric chemistry processes. See also Carbon dioxide
addition, parties may use other metrics to report supplemental
removal (CDR), Net negative greenhouse gas emissions, Net zero CO2
information on aggregate emissions and removals of GHGs.
emissions and Net zero greenhouse gas emissions.
See also Net zero CO2 emissions, Greenhouse gas emission metric
Net negative greenhouse gas emissions A situation of
and Greenhouse gas neutrality).
net negative greenhouse gas emissions is achieved when metric-
weighted anthropogenic greenhouse gas (GHG) removals exceed Nitrogen deposition Nitrogen deposition is defined as the
metric-weighted anthropogenic GHG emissions. Where multiple nitrogen transferred from the atmosphere to the Earth’s surface by
GHG are involved, the quantification of net emissions depends on the processes of wet deposition and dry deposition.
the metric chosen to compare emissions of different gases (such as
Nitrous oxide (N2O) The main anthropogenic source of N2O,
global warming potential, global temperature change potential, and
a greenhouse gas (GHG), is agriculture (soil and animal manure
others, as well as the chosen time horizon). See also Net zero CO2
management), but important contributions also come from sewage
emissions, Net zero greenhouse gas emissions, Negative greenhouse
treatment, fossil fuel combustion, and chemical industrial processes.
gas emissions, Carbon dioxide removal (CDR) and Greenhouse gas
N2O is also produced naturally from a wide variety of biological
emission metric.
sources in soil and water, particularly microbial action in wet
Net primary production (NPP) See Primary production. tropical forests.
Net zero CO2 emissions Condition in which anthropogenic Non-CO2 emissions and radiative forcing Non-CO2 emissions
carbon dioxide (CO2) emissions are balanced by anthropogenic CO2 included in this report are all anthropogenic emissions other than
removals over a specified period. carbon dioxide (CO2) that result in radiative forcing. These include
short-lived climate forcers, such as methane (CH4 ), some fluorinated
Note: Carbon neutrality and net zero CO2 emissions are overlapping
gases, ozone (O3 ) precursors, aerosols or aerosol precursors, such as
concepts. The concepts can be applied at global or sub-global scales
black carbon and sulphur dioxide, respectively, as well as long-lived
(e.g., regional, national and sub-national). At a global scale, the
greenhouse gases, such as nitrous oxide (N2O) or other fluorinated
terms carbon neutrality and net zero CO2 emissions are equivalent.
gases. The radiative forcing associated with non-CO2 emissions and
At sub-global scales, net zero CO2 emissions is generally applied
changes in surface albedo (e.g., resulting from land-use change) is
to emissions and removals under direct control or territorial
referred to as non-CO2 radiative forcing.
responsibility of the reporting entity, while carbon neutrality
generally includes emissions and removals within and beyond the Non-linearity A process is called non-linear when there is no
direct control or territorial responsibility of the reporting entity. simple proportional relation between cause and effect. The climate
Accounting rules specified by GHG programmes or schemes can system contains many such non-linear processes, resulting in
have a significant influence on the quantification of relevant CO2 a system with potentially very complex behaviour. Such complexity
emissions and removals. may lead to abrupt climate change and tipping points.
See also Net zero greenhouse gas emissions and Carbon neutrality. Non-methane volatile organic compounds (NMVOCs) See
Volatile organic compounds (VOCs).
Net zero greenhouse gas emissions Condition in which
metric-weighted anthropogenic greenhouse gas (GHG) emissions Non-overshoot pathways See Pathways.
are balanced by metric-weighted anthropogenic GHG removals over
North American monsoon (NAmerM) See Global monsoon.
a specified period. The quantification of net zero GHG emissions
depends on the GHG emission metric chosen to compare emissions Northern Annular Mode (NAM) See Annular modes.
and removals of different gases, as well as the time horizon chosen
North Atlantic Oscillation (NAO) The leading mode of large-
for that metric.
scale atmospheric variability in the North Atlantic basin characterized
Note 1: GHG neutrality and net zero GHG emissions are overlapping by alternating (see-saw) variations in sea level pressure or
concepts. The concept of net zero GHG emissions can be applied geopotential height between the Azores High in the subtropics and
at global or sub-global scales (e.g., regional, national and sub- the Icelandic Low in the mid- to high latitudes, with some northward
AVII national). At a global scale, the terms GHG neutrality and net zero extension deep into the Arctic. It is associated with fluctuations in the
GHG emissions are equivalent. At sub-global scales, net zero GHG strength and latitudinal position of the main westerly winds across
244
Glossary Annex VII
a vast North Atlantic–Europe domain, and thus with fluctuations in from such an experiment as the ratio of the rate of increase of ocean
the embedded extratropical cyclones and associated frontal systems heat content to the surface temperature change.
leading to strong teleconnection over the entire North Atlantic
Ocean stratification See Stratification.
adjacent continents. The positive and negative phases of the NAO
show similar characteristics described for the Northern Annular Mode Orbital forcing Orbital forcing is the influence of slow, systematic
(NAM). See Section AIV.2.1 in Annex IV of the AR6 WGI report. and predictable changes in orbital parameters (eccentricity, obliquity
and precession of the equinox) on incoming solar radiation
Northern polar vortex See Stratospheric polar vortex.
(insolation), especially its latitudinal and seasonal distribution. It is
Ocean The interconnected body of saline water that covers 71% of an external forcing and a key driver of glacial–interglacial cycles.
the Earth’s surface, contains 97% of the Earth’s water and provides
Organic aerosol Component of the aerosol that consists of
99% of the Earth’s biologically habitable space. It includes the
organic compounds, mainly carbon, hydrogen, oxygen and lesser
Arctic, Atlantic, Indian, Pacific and Southern Oceans, as well as their
amounts of other elements.
marginal seas and coastal waters.
Outgoing longwave radiation Net outgoing radiation in the
Ocean acidification (OA) A reduction in the pH of the ocean,
infrared part of the spectrum at the top of the atmosphere.
accompanied by other chemical changes (primarily in the levels of
carbonate and bicarbonate ions), over an extended period, typically Outlet glacier See Glacier.
decades or longer, which is caused primarily by uptake of carbon
Overshoot pathways See Pathways.
dioxide (CO2) from the atmosphere, but can also be caused by other
chemical additions or subtractions from the ocean. Anthropogenic Oxygen minimum zone (OMZ) The midwater layer (200–
OA refers to the component of pH reduction that is caused by human 1000 m) in the open ocean in which oxygen saturation is the lowest
activity (IPCC, 2011, p. 37). in the ocean. The degree of oxygen depletion depends on the largely
bacterial consumption of organic matter, and the distribution of the
Ocean alkalinization/Ocean alkalinity enhancement A
OMZs is influenced by large-scale ocean circulation. In coastal oceans,
proposed carbon dioxide removal (CDR) method that involves
OMZs extend to the shelves and may also affect benthic ecosystems.
deposition of alkaline minerals or their dissociation products at the
ocean surface. This increases surface total alkalinity, and may thus Ozone (O3) The triatomic form of oxygen, and a gaseous
increase ocean carbon dioxide (CO2) uptake and ameliorate surface atmospheric constituent. In the troposphere, O3 is created both
ocean acidification. See also Anthropogenic removals. naturally and by photochemical reactions involving gases resulting
from human activities (e.g., smog). Tropospheric O3 acts as
Ocean carbon cycle The ocean carbon cycle is the set of processes
a greenhouse gas (GHG). In the stratosphere, O3 is created by the
that exchange carbon between various pools within the ocean, as
interaction between solar ultraviolet radiation and molecular oxygen
well as between the atmosphere, Earth’s interior, cryosphere, and the
(O2). Stratospheric O3 plays a dominant role in the stratospheric
sea-floor. See also Carbon cycle.
radiative balance. Its concentration is highest in the ozone layer. See
Ocean deoxygenation The loss of oxygen in the ocean. It results also Ground-level ozone, Ozone hole, Ozone-depleting substances
from ocean warming, which reduces oxygen solubility and increases (ODSs), Ozonesonde and Short-lived climate forcers (SLCFs).
oxygen consumption and stratification, thereby reducing the mixing
Ozone layer A layer of Earth’s stratosphere that absorbs most
of oxygen into the ocean interior. Deoxygenation can also be
of the Sun’s ultraviolet radiation. It contains high concentrations of
exacerbated by the addition of excess nutrients in the coastal zone.
ozone (O3 ) in relation to other parts of the atmosphere, although still
Ocean dynamic sea level change See Sea level change (sea small in relation to other gases in the stratosphere. The ozone layer
level rise/sea level fall). contains less than 10 parts per million of ozone, while the average
ozone concentration in Earth’s atmosphere as a whole is about
Ocean fertilization A proposed carbon dioxide removal (CDR)
0.3 parts per million. The ozone layer is mainly found in the lower
method that relies on the deliberate increase of nutrient supply to
portion of the stratosphere, from approximately 15 to 35 kilometres
the near-surface ocean with the aim of sequestering additional CO2
(9.3 to 21.7 miles) above Earth, although its thickness varies
from the atmosphere through biological production. Methods include
seasonally and geographically. See also Ozone hole and Ozone-
direct addition of micro-nutrients or macro-nutrients. To be successful,
depleting substances (ODSs).
the additional carbon needs to reach the deep ocean where it has the
potential to be sequestered on climatically relevant time scales. See Ozone-depleting substances (ODSs) Man-made gases that
also Anthropogenic removals and Carbon dioxide removal (CDR). destroy ozone (O3 ) once they reach the ozone layer in the stratosphere.
Ozone-depleting substances include: chlorofluorocarbons (CFCs),
Ocean heat uptake efficiency This is a measure (W m–2 °C–1) of
hydrochlorofluorocarbons (HCFCs), hydrobromofluorocarbons
the rate at which heat storage by the global ocean increases as global
(HBFCs), halons, methyl bromide, carbon tetrachloride and methyl
surface temperature rises. It is a useful parameter for climate change
chloroform. They are used as refrigerants in commercial, home and
simulations in which the radiative forcing is changing monotonically,
vehicle air conditioners and refrigerators, foam blowing agents,
when it can be compared with the climate feedback parameter to
components in electrical equipment, industrial solvents, solvents
gauge the relative importance of radiative response and ocean heat
for cleaning (including dry cleaning), aerosol spray propellants and
uptake in determining the rate of climate change. It can be estimated AVII
fumigants. See also Ozone layer, Ozone (O3 ) and Stratospheric ozone.
245
Annex VII Glossary
Ozonesonde An ozonesonde is a radiosonde measuring ozone low pressure anomalies stretching from the subtropical west Pacific
(O3 ) concentrations. The radiosonde is usually carried on a weather to the east coast of North America.
balloon and transmits measured quantities by radio to a ground-
Palaeocene–Eocene Thermal Maximum (PETM) The PETM is
based receiver.
a transient event that occurred between 55.9 and 55.7 million years
Pacific Decadal Oscillation (PDO) See Pacific Decadal ago. Continental positions at this time were somewhat different to
Variability (PDV). present due to tectonic plate movements. Geological data indicate
that the PETM was characterised by a warming (global mean surface
Pacific Decadal Variability (PDV) Coupled decadal-to-inter-
temperature rose to about 4°C–7°C warmer than the preceding mean
decadal variability of the atmospheric circulation and underlying
state), and an increase in atmospheric CO2 (from about 900 to about
ocean that is typically observed over the entire Pacific Basin
2000 ppmv). In addition, ocean pH and oxygen content decreased;
beyond the El Niño–Southern Oscillation (ENSO) time scale. In the
many deep-sea species went extinct and tropical coral reefs diminished.
AR6 WGI report, PDV encapsulates the Pacific Decadal Oscillation
(PDO), the South Pacific Decadal Oscillation (SPDO), tropical Pacific Paleoclimate Climate during periods prior to the development
decadal variability (also called decadal ENSO), and the Inter-decadal of measuring instruments, including historic and geologic time, for
Pacific Oscillation (IPO). Typically, the positive phase of the PDV is which only proxy climate records are available.
characterized by anomalously high sea surface temperatures in
Parameterization In climate models, this term refers to the
the central-eastern tropical Pacific that extend to the extratropical
technique of representing processes that cannot be explicitly
North and South Pacific along the American coasts, encircled to the
resolved at the spatial or temporal resolution of the model (sub-grid
west by cold sea surface anomalies in the mid-latitude North and
scale processes) by relationships between model-resolved larger-
South Pacific. The negative phase is accompanied by sea surface
scale variables and the area- or time-averaged effect of such sub-grid
temperature anomalies of the opposite sign. Those sea surface
scale processes.
temperature anomalies are linked to anomalies in atmospheric and
oceanic circulation throughout the whole Pacific Basin. The PDV is Pathways The temporal evolution of natural and/or human
associated with decadal modulations in the relative occurrence of systems towards a future state. Pathway concepts range from sets
El Niño and La Niña. See Section AIV.2.6 in Annex IV of the AR6 of quantitative and qualitative scenarios or narratives of potential
WGI report. futures to solution-oriented decision-making processes to achieve
desirable societal goals. Pathway approaches typically focus on
Inter-decadal Pacific Oscillation (IPO)
biophysical, techno-economic, and/or socio-behavioural trajectories
An equatorially symmetric pattern of sea surface temperature
and involve various dynamics, goals, and actors across different
variability at decadal-to-inter-decadal time scales. While the Pacific
scales. See also Scenario and Scenario storyline (under Storyline).
Decadal Oscillation (PDO) and its South Pacific counterpart, the
South Pacific Decadal Oscillation (SPDO), are considered as physically 1.5°C pathway
distinct modes, the tropical Pacific decadal–inter-decadal variability A pathway of emissions of greenhouse gases and other climate
can drive both the PDO and SPDO, forming the IPO as a synchronized forcers that provides an approximately one-in-two to two-in-three
pan-Pacific variability. Its spatial pattern of sea surface temperature chance, given current knowledge of the climate response, of global
anomalies is similar to that of the El Niño–Southern Oscillation warming either remaining below 1.5°C or returning to 1.5°C by
(ENSO), but with a broader meridional extent in the tropical signal around 2100 following an overshoot.
and more weights in the extratropics compared to the tropics. In
Emission pathways
the AR6 WGI report, it is encapsulated within the definition and
Modelled trajectories of global anthropogenic emissions over the
description of Pacific Decadal Variability (PDV). See also Section
21st century.
AIV.2.6 in Annex IV of the AR6 WGI report.
Mitigation pathways
Pacific Decadal Oscillation (PDO)
A temporal evolution of a set of mitigation scenario features, such
The leading mode of variability obtained from decomposition in
as greenhouse gas (GHG) emissions and socio-economic development.
empirical orthogonal function of sea surface temperature over the
North Pacific north of 20°N, and characterized by a strong decadal Non-overshoot pathways
component. The positive phase of the PDO features a dipole of sea Pathways that stay below a specified concentration, forcing, or global
surface temperature anomalies in the North Pacific, with a cold lobe warming level during a specified period of time (e.g., until 2100).
near the centre of the basin and extending westward along the
Kuroshio, encircled by warmer conditions along the coast of North Overshoot pathways
America and in the subtropics. A positive PDO is accompanied by Pathways that first exceed a specified concentration, forcing, or global
an intensified Aleutian Low and an associated cyclonic circulation warming level, and then return to or below that level again before
enhancement leading to teleconnections over the continents adjacent the end of a specified period of time (e.g., before 2100). Sometimes
to the North Pacific. In the AR6 WGI report, the PDO is encapsulated the magnitude and likelihood of the overshoot is also characterized.
within the definition and description of Pacific Decadal Variability The overshoot duration can vary from one pathway to the next,
(PDV). See also Section AIV.2.6 in Annex IV of the AR6 WGI report. but in most overshoot pathways in the literature and referred to as
overshoot pathways in the AR6, the overshoot occurs over a period
AVII Pacific-North American (PNA) pattern An atmospheric large- of at least one decade and up to several decades.
scale wave pattern featuring a sequence of tropospheric high and
246
Glossary Annex VII
Representative Concentration Pathways (RCPs) within an integrated assessment model. Those SSP scenarios are
Scenarios that include time series of emissions and concentrations bare of climate policy assumption, but in combination with so-
of the full suite of greenhouse gases (GHGs) and aerosols and called shared policy assumptions (SPAs), various approximate
chemically active gases, as well as land use/land cover (Moss et radiative forcing levels of 1.9, 2.6, …, or 8.5 W m–2 are reached by
al., 2010). The word representative signifies that each RCP provides the end of the century, respectively.
only one of many possible scenarios that would lead to the specific
Pattern scaling Techniques used to represent the spatial
radiative forcing characteristics. The term pathway emphasises
variations in climate at a given increase in global mean surface air
that not only the long-term concentration levels are of interest, but
temperature (GSAT) are referred to as ‘pattern scaling’.
also the trajectory taken over time to reach that outcome (Moss et
al., 2010). Peat Soft, porous or compressed, sedimentary deposit of which
a substantial portion is partly decomposed plant material with high
RCPs usually refer to the portion of the concentration pathway
water content in the natural state (up to about 90%).
extending up to 2100, for which integrated assessment models
produced corresponding emission scenarios. Extended concentration Peatlands Peatland is a land where soils are dominated by peat.
pathways describe extensions of the RCPs from 2100 to 2300
Percentile A partition value in a population distribution that
that were calculated using simple rules generated by stakeholder
a given percentage of the data values are below or equal to. The 50th
consultations, and do not represent fully consistent scenarios. Four
percentile corresponds to the median of the population. Percentiles
RCPs produced from integrated assessment models were selected
are often used to estimate the extremes of a distribution. For example,
from the published literature and used in the Fifth IPCC Assessment,
the 90th (10th) percentile may be used to refer to the threshold for
and are also used in this Assessment for comparison, spanning
the upper (lower) extremes.
the range from approximately below 2°C warming to high (>4°C)
warming best-estimates by the end of the 21st century: RCP2.6, Permafrost Ground (soil or rock, and included ice and organic
RCP4.5 and RCP6.0 and RCP8.5. material) that remains at or below 0°C for at least two consecutive
years (Harris et al., 1988). Note that permafrost is defined via
• RCP2.6: One pathway where radiative forcing peaks at
temperature rather than ice content and, in some instances, may
approximately 3 W m–2 and then declines to be limited at 2.6 W
be ice-free.
m–2 in 2100 (the corresponding Extended Concentration Pathway,
or ECP, has constant emissions after 2100). Near-surface permafrost
Permafrost within about 3–4 m of the ground surface. The depth
• RCP4.5 and RCP6.0: Two intermediate stabilization pathways
is not precise, but describes what commonly is highly relevant for
in which radiative forcing is limited at approximately 4.5 W m–2
people and ecosystems. Deeper permafrost is often progressively
and 6.0 W m–2 in 2100 (the corresponding ECPs have constant
less ice-rich and responds more slowly to warming than near-surface
concentrations after 2150).
permafrost. The presence or absence of near-surface permafrost is
• RCP8.5: One high pathway which leads to >8.5 W m–2 in 2100 not the only significant metric of permafrost change, and deeper
(the corresponding ECP has constant emissions after 2100 until permafrost may persist when near-surface permafrost is absent.
2150 and constant concentrations after 2250).
Permafrost degradation
See also Coupled Model Intercomparison Project (CMIP) and Shared Decrease in the thickness and/or areal extent of permafrost.
Socio-economic Pathways (SSPs) (under Pathways).
Permafrost thaw
Shared Socio-economic Pathways (SSPs) Progressive loss of ground ice in permafrost, usually due to input of
Shared Socio-economic Pathways (SSPs) have been developed heat. Thaw can occur over decades to centuries over the entire depth
to complement the Representative Concentration Pathways of permafrost ground, with impacts occurring while thaw progresses.
(RCPs). By design, the RCP emission and concentration pathways During thaw, temperature fluctuations are subdued because energy
were stripped of their association with a certain socio-economic is transferred by phase change between ice and water. After the
development. Different levels of emissions and climate change transition from permafrost to non-permafrost, ground can be
along the dimension of the RCPs can hence be explored against described as thawed.
the backdrop of different socio-economic development pathways
Perturbed parameter ensemble See Climate simulation
(SSPs) on the other dimension in a matrix. This integrative SSP-RCP
ensemble.
framework is now widely used in the climate impact and policy
analysis literature, where climate projections obtained under the pH A dimensionless measure of the acidity of a dilute solution
RCP scenarios are analysed against the backdrop of various SSPs. (e.g., seawater) based on the activity, or effective concentration, of
As several emissions updates were due, a new set of emissions hydrogen ions (H+) in the solution. pH is measured on a logarithmic
scenarios was developed in conjunction with the SSPs. Hence, the scale where pH = –log10(H+). Thus, a pH decrease of 1 unit corresponds
abbreviation SSP is now used for two things: On the one hand to a 10-fold increase in the acidity, or the activity of H+.
SSP1, SSP2, …, SSP5 are used to denote the five socio-economic
scenario families. On the other hand, the abbreviations SSP1-1.9, Phenology The relationship between biological phenomena that
SSP1-2.6, …, SSP5-8.5 are used to denote the newly developed recur periodically (e.g., development stages, migration) and climate
and seasonal changes. AVII
emissions scenarios that are the result of an SSP implementation
247
Annex VII Glossary
Photosynthesis The production of carbohydrates in plants, algae expressed in terms of the height of the water if completely condensed
and some bacteria using the energy of light. Carbon dioxide (CO2) is and collected in a vessel of the same unit cross section.
used as the carbon source.
Precursors Atmospheric compounds that are not greenhouse
Physical climate storyline See Storyline. gases (GHGs) or aerosols, but that have an effect on GHG or aerosol
concentrations by taking part in physical or chemical processes
Piacenzian warm period See Pliocene.
regulating their production or destruction rates.
Plankton Free-floating organisms living in the upper layers
Predictability The extent to which future states of a system
of aquatic systems. Their distribution and migration are primarily
may be predicted based on knowledge of current and past states
determined by water currents. A distinction is made between
of the system. Because knowledge of the climate system’s past and
phytoplankton, which depend on photosynthesis for their energy
current states is generally imperfect, as are the models that utilize
supply, and zooplankton, which feed on phytoplankton, other
this knowledge to produce a climate prediction, and because the
zooplankton, and bacterioplankton.
climate system is inherently non-linear and chaotic, predictability of
Plant evaporative stress Plant evaporative stress in both the climate system is inherently limited. Even with arbitrarily accurate
crops and natural vegetation can result from the combination of models and observations, there may still be limits to the predictability
a high atmospheric evaporative demand and limited available water of such a non-linear system (AMS, 2021). See also Climate prediction
to supply this demand by means of evapotranspiration, further and Prediction quality/skill.
enhancing agricultural and ecological drought.
Prediction quality/skill Measures of the success of a prediction
Pleistocene The Pleistocene Epoch is the earlier of two epochs in against observationally based information. No single measure can
the Quaternary System, extending from 2.59 Ma to the beginning of summarize all aspects of forecast quality, and a suite of metrics is
the Holocene at approximately 11.7 ka. considered. Metrics will differ for forecasts given in deterministic and
probabilistic form. See also Climate prediction and Predictability.
Pliocene The Pliocene Epoch is the more recent of two epochs of
the Neogene Period within the Cenozoic Era. It extends from 5.33 Ma Primary production The synthesis of organic compounds
to the beginning of the Pleistocene Epoch at 2.59 Ma. The Neogene by plants and microbes, on land or in the ocean, primarily by
Period precedes the current geological period, the Quaternary Period, photosynthesis using light and carbon dioxide (CO2) as sources
which is one of several ice ages that have occurred during Earth’s of energy and carbon respectively. It can also occur through
geological history. It encompasses the mid-Pliocene warm period chemosynthesis, using chemical energy, for example, in deep
(MPWP), also known as the Piacenzian warm period, which occurred sea vents.
from approximately 3.3 to 3.0 Ma. The MPWP, in turn, encompasses
Gross primary production (GPP)
the interglacial episode, marine isotope stage (MIS) KM5c, which
The total amount of carbon fixed by photosynthesis over a specified
peaked at 3.205 Ma, when orbital forcing was similar to modern
time period.
(Haywood et al., 2016).
Net primary production (NPP)
Polar amplification Polar amplification describes the
The amount of carbon fixed by photosynthesis minus the amount lost
phenomenon where surface temperature change at high latitudes
by respiration over a specified time period.
exceeds the global average surface temperature change. The terms
Arctic amplification or Antarctic amplification are used when Probability density function (PDF) A probability density
describing the phenomenon occurring at one of the poles. function is a function that indicates the relative chances of occurrence
of different outcomes of a variable. The function integrates to unity
Pollen analysis A technique of both relative dating and
over the domain for which it is defined and has the property that
environmental reconstruction, consisting of the identification
the integral over a sub-domain equals the probability that the
and counting of pollen types preserved in peat, lake sediments and
outcome of the variable lies within that sub-domain. For example,
other deposits.
the probability that a temperature anomaly defined in a particular
Pool, carbon and nitrogen A reservoir in the Earth system way is greater than zero is obtained from its PDF by integrating
where elements, such as carbon and nitrogen, reside in various the PDF over all possible temperature anomalies greater than zero.
chemical forms for a period of time. See also Reservoir, Sequestration, Probability density functions that describe two or more variables
Sequestration potential, Sink, Source and Uptake. simultaneously are similarly defined.
Post-glacial period See Holocene. Process-based model Theoretical concepts and computational
methods that represent and simulate the behaviour of real-world
Potential evapotranspiration See Evapotranspiration.
systems derived from a set of functional components and their
Pre-industrial (period) The multi-century period prior to the interactions with each other and the system environment, through
onset of large-scale industrial activity around 1750. The reference physical and mechanistic processes occurring over time.
period 1850–1900 is used to approximate pre-industrial global mean
Projection A potential future evolution of a quantity or set of
surface temperature (GMST). See also Industrial revolution.
quantities, often computed with the aid of a model. Unlike predictions,
AVII Precipitable water The total amount of atmospheric water projections are conditional on assumptions concerning, for example,
vapour in a vertical column of unit cross-sectional area. It is commonly future socio-economic and technological developments that may
248
Glossary Annex VII
or may not be realized. See also Climate projection, Pathways system that occur in operational analyses. Although continuity is
and Scenario. improved, global reanalyses still suffer from changing coverage and
biases in the observing systems.
Proxy A proxy climate indicator is any biophysical property of
materials formed during the past that is interpreted to represent Reasons for concern (RFCs) Elements of a classification
some combination of climate-related variations back in time. framework, first developed in the IPCC Third Assessment Report,
Climate-related data derived in this way are referred to as proxy data, which aims to facilitate judgements about what level of climate
and time series of proxy data are proxy records. Examples of proxy change may be dangerous (in the language of Article 2 of the
types include pollen assemblages, tree ring widths, speleothem and UNFCCC; UNFCCC, 1992) by aggregating risks from various sectors,
coral geochemistry, and various data derived from marine sediments considering hazards, exposures, vulnerabilities, capacities to adapt,
and glacier ice. Proxy data can be calibrated to provide quantitative and the resulting impacts.
climate information.
Reconstruction (of climate variable) Approach to
Proxy records See Proxy. reconstructing the past temporal and spatial characteristics
of a climate variable from predictors. The predictors can be
Quasi-Biennial Oscillation (QBO) A near-periodic oscillation of
instrumental data if the reconstruction is used to infill missing data
the equatorial zonal wind between easterlies and westerlies in the
or proxy data if it is used to develop paleoclimate reconstructions.
tropical stratosphere with a mean period of around 28 months. The
Various techniques have been developed for this purpose: linear
alternating wind maxima descend from the base of the mesosphere
multivariate regression-based methods and non-linear Bayesian
down to the tropopause and are driven by wave energy that
and analogue methods.
propagates up from the troposphere.
Reference period A time period of interest, or a period over
Quaternary The Quaternary Period is the last of three periods that
which some relevant statistics are calculated. A reference period can
make up the Cenozoic Era (66 Ma to present), extending from 2.58 Ma
be used as a baseline period or as a comparison to a baseline period.
to the present, and includes the Pleistocene and Holocene Epochs.
Baseline period
Radiative forcing The change in the net, downward minus
A time period against which differences are calculated (e.g., expressed
upward, radiative flux (expressed in W m–2) due to a change in
as anomalies relative to a baseline).
an external driver of climate change, such as a change in the
concentration of carbon dioxide (CO2), the concentration of volcanic Reference scenario See Scenario.
aerosols or the output of the Sun. The stratospherically adjusted
Reforestation Conversion to forest of land that has previously
radiative forcing is computed with all tropospheric properties held
contained forests but that has been converted to some other use.
fixed at their unperturbed values, and after allowing for stratospheric
[Note: For a discussion of the term forest and related terms such
temperatures, if perturbed, to readjust to radiative-dynamical
as afforestation, reforestation and deforestation, see the 2006
equilibrium. Radiative forcing is called instantaneous if no change in
IPCC Guidelines for National Greenhouse Gas Inventories and
stratospheric temperature is accounted for. The radiative forcing once
their 2019 Refinement, and information provided by the United
both stratospheric and tropospheric adjustments are accounted for is
Nations Framework Convention on Climate Change (IPCC, 2006,
termed the effective radiative forcing.
2019; UNFCCC, 2021a, b).] See also Afforestation, Deforestation,
Radiative response (of the climate system) The net top-of- Anthropogenic removals and Carbon dioxide removal (CDR).
atmosphere radiative flux that opposes a change in radiative forcing
Region Land and/or ocean area characterized by specific
as a result of climate feedbacks. Typical units: W m–2. See also Earth’s
geographical and/or climatological features. The climate of a region
energy budget and Climate feedback parameter.
emerges from a multi-scale combination of its own features, remote
Rapid dynamical change (of glaciers or ice sheets) Changes influences from other regions, and global climate conditions.
in glacier or ice sheet mass controlled by changes in flow speed and
Regional climate model (RCM) A climate model at higher
discharge rather than by accumulation or ablation. This can result in
resolution over a limited area. Such models are used in downscaling
a rate of mass change larger than that due to any imbalance between
global climate results over specific regional domains.
accumulation and ablation. Rapid dynamical change may be initiated
by a climatic trigger, such as incursion of warm ocean water beneath Regional sea level change See Sea level change (sea level rise/
an ice shelf, or thinning of a grounded tide-water terminus, which sea level fall).
may lead to reactions within the glacier system that may result in
Relative humidity The ratio of actual water vapour pressure to
rapid ice loss.
that at saturation with respect to liquid water or ice at the same
Reanalysis Reanalyses are created by processing past temperature. See also Specific humidity.
meteorological or oceanographic data using fixed state-of-the-
Relative sea level (RSL) change See Sea level change (sea level
art weather forecasting or ocean circulation models with data
rise/sea level fall).
assimilation techniques. They are used to provide estimates of
variables such as historical atmospheric temperature and wind or Remaining carbon budget See Carbon budget.
oceanographic temperature and currents, and other quantities. Using
Representative Concentration Pathways (RCPs) See AVII
fixed data assimilation avoids effects from the changing analysis
Pathways.
249
Annex VII Glossary
Reservoir A component or components of the climate system and vulnerability of the affected human or ecological system to the
where a greenhouse gas (GHG) or a precursor of a greenhouse gas is hazards. Hazards, exposure and vulnerability may each be subject
stored (UNFCCC Article 1.7 (UNFCCC, 1992)). See also Pool, carbon to uncertainty in terms of magnitude and likelihood of occurrence,
and nitrogen, Sequestration, Sequestration potential, Sink, Source and each may change over time and space due to socio-economic
and Uptake. changes and human decision-making (see also risk management,
adaptation and mitigation).
Resilience The capacity of interconnected social, economic
and ecological systems to cope with a hazardous event, trend or In the context of climate change responses, risks result from the
disturbance, responding or reorganizing in ways that maintain their potential for such responses not achieving the intended objective(s),
essential function, identity and structure. Resilience is a positive or from potential trade-offs with, or negative side-effects on, other
attribute when it maintains capacity for adaptation, learning and/ societal objectives, such as the Sustainable Development Goals
or transformation (Arctic Council, 2016). See also Hazard, Risk (SDGs) (see also risk trade-off). Risks can arise, for example, from
and Vulnerability. uncertainty in implementation, effectiveness or outcomes of climate
policy, climate-related investments, technology development or
Resolution In climate models, this term refers to the physical
adoption, and system transitions. See also Hazard and Impacts
distance (metres or degrees) between each point on the grid used to
(consequences, outcomes).
compute the equations. Temporal resolution refers to the time step
or time elapsed between each model computation of the equations. Risk assessment The qualitative and/or quantitative scientific
estimation of risks. See also Risk management and Risk perception.
Respiration The process whereby living organisms convert
organic matter to carbon dioxide (CO2), releasing energy and Risk framework A common framework for describing and
consuming molecular oxygen. assessing risk across all three Working Groups is adopted to promote
clear and consistent communication of risks and to better inform risk
Response time or adjustment time In the context of climate
assessment and decision-making related to climate change.
variations, the response time or adjustment time is the time needed
for the climate system or its components to re-equilibrate to a new Risk management Plans, actions, strategies or policies to reduce
state, following a forcing resulting from external processes. It is the likelihood and/or magnitude of adverse potential consequences,
very different for various components of the climate system. The based on assessed or perceived risks. See also Risk assessment
response time of the troposphere is relatively short, from days to and Risk perception.
weeks, whereas the stratosphere reaches equilibrium on a time
Risk perception The subjective judgement that people make
scale of typically a few months. Due to their large heat capacity,
about the characteristics and severity of a risk. See also Risk
the oceans have a much longer response time: typically decades,
assessment and Risk management.
but up to centuries or millennia. The response time of the strongly
coupled surface–troposphere system is, therefore, slow compared Risk trade-off The change in the portfolio of risks that occurs when
to that of the stratosphere, and mainly determined by the oceans. a countervailing risk is generated (knowingly or inadvertently) by an
The biosphere may respond quickly (e.g., to droughts), but also very intervention to reduce the target risk (Wiener and Graham, 2009).
slowly to imposed changes.
River discharge See Streamflow.
In the context of lifetimes, response time or adjustment time (Ta) is
Rock glacier A debris landform (mass of rock fragments and finer
the time scale characterizing the decay of an instantaneous pulse
material that contains either an ice core or an ice-cemented matrix)
input into the reservoir. See Response time or adjustment time (Ta )
generated by a former or current gravity-driven creep of permafrost
under Lifetime.
in mountain slopes (Harris et al., 1988; Giardino et al., 2011; IPA-RG,
Return period An estimate of the average time interval between 2020). It is detectable in the landscape due to the occurrence of (i)
occurrences of an event (e.g., flood or extreme rainfall) of (or below/ a steep slope delimiting the terminal part, (ii) generally well-defined
above) a defined size or intensity. lateral margins in a continuation of the front, and (iii) transversal
or longitudinal ridges and furrows (ridge and furrow topography).
Return value The highest (or, alternatively, lowest) value of
These are geomorphological indicators of the occurrence of
a given variable, on average occurring once in a given period of time
permafrost conditions. Although it is an ice storage feature, it is not
(e.g., in 10 years).
a type of glacier since it does not originate at the surface by the
Risk The potential for adverse consequences for human or recrystallization of snow.
ecological systems, recognizing the diversity of values and objectives
Runoff The flow of water over the surface or through the
associated with such systems. In the context of climate change,
subsurface, which typically originates from the part of liquid
risks can arise from potential impacts of climate change as well as
precipitation and/or snow/ice melt that does not evaporate, transpire
human responses to climate change. Relevant adverse consequences
or refreeze, and returns to water bodies.
include those on lives, livelihoods, health and well-being, economic,
social and cultural assets and investments, infrastructure, services Sampling uncertainty See Uncertainty.
(including ecosystem services), ecosystems and species.
Scenario A plausible description of how the future may develop
AVII In the context of climate change impacts, risks result from dynamic based on a coherent and internally consistent set of assumptions
interactions between climate-related hazards with the exposure about key driving forces (e.g., rate of technological change (TC),
250
Glossary Annex VII
prices) and relationships. Note that scenarios are neither predictions assuming no anthropogenic greenhouse gas (GHG) emissions (climate
nor forecasts, but are used to provide a view of the implications of change attribution) or no climate change (impact attribution).]
developments and actions. See also Pathways and Scenario storyline
Socio-economic scenario
(under Storyline).
A scenario that describes a plausible future in terms of population,
Baseline scenario gross domestic product (GDP), and other socio-economic factors
See Reference scenario (under Scenario). relevant to understanding the implications of climate change.
Concentrations scenario Scenario storyline See Storyline.
A plausible representation of the future development of atmospheric
Sea ice Ice found at the sea surface that has originated from
concentrations of substances that are radiatively active (e.g.,
the freezing of seawater. Sea ice may be discontinuous pieces (ice
greenhouse gases (GHGs), aerosols, tropospheric ozone), plus
floes) moved on the ocean surface by wind and currents (pack ice),
human-induced land–cover changes that can be radiatively active
or a motionless sheet attached to the coast (land-fast ice). Sea ice
via albedo changes, and often used as input to a climate model to
concentration is the fraction of the ocean covered by ice. Sea ice less
compute climate projections.
than one year old is called first-year ice. Perennial ice is sea ice that
Emissions scenario survives at least one summer. It may be subdivided into second-year
A plausible representation of the future development of emissions of ice and multi-year ice, where multi-year ice has survived at least
substances that are radiatively active (e.g., greenhouse gases (GHGs) two summers.
or aerosols), plus human-induced land-cover changes that can be
Sea ice area (SIA)
radiatively active via albedo changes, based on a coherent and
Sea ice area is the area covered by sea ice. In contrast to sea ice
internally consistent set of assumptions about driving forces (such
extent, it is a linear measure of sea ice coverage that does not depend
as demographic and socio-economic development, technological
on grid resolution.
change, energy and land use) and their key relationships.
Concentration scenarios, derived from emission scenarios, are often Sea ice concentration
used as input to a climate model to compute climate projections. Sea ice concentration is the fraction of the ocean covered by ice.
Mitigation scenario Sea ice extent (SIE) Sea ice extent is calculated for gridded data
A plausible description of the future that describes how the (studied) products as the total area of all grid cells with sea ice concentration
system responds to the implementation of mitigation policies and above a given threshold, usually 15 %. It hence is a grid-dependent,
measures. non-linear measure of sea ice coverage.
Reference scenario Sea level change (sea level rise/sea level fall) Change to
Scenario used as starting or reference point for a comparison between the height of sea level, both globally and locally (relative sea level
two or more scenarios. change) at seasonal, annual, or longer time scales due to (i) a change
in ocean volume as a result of a change in the mass of water in the
[Note 1: In many types of climate change research, reference
ocean (e.g., due to melt of glaciers and ice sheets), (ii) changes in
scenarios reflect specific assumptions about patterns of socio-
ocean volume as a result of changes in ocean water density (e.g.,
economic development and may represent futures that assume
expansion under warmer conditions), (iii) changes in the shape of the
no climate policies or specified climate policies, for example those
ocean basins and changes in the Earth’s gravitational and rotational
in place or planned at the time a study is carried out. Reference
fields, and (iv) local subsidence or uplift of the land. Global mean sea
scenarios may also represent futures with limited or no climate
level (GMSL) change resulting from change in the mass of the ocean
impacts or adaptation, to serve as a point of comparison for futures
is called barystatic. The amount of barystatic sea level change due
with impacts and adaptation. These are also referred to as baseline
to the addition or removal of a mass of water is called its sea level
scenarios in the literature.
equivalent (SLE). Sea level changes, both globally and locally, resulting
Note 2: Reference scenarios can also be climate policy or impact from changes in water density are called steric. Density changes
scenarios, which in that case are taken as a point of comparison induced by temperature changes only are called thermosteric, while
to explore the implications of other features, for example, of delay, density changes induced by salinity changes are called halosteric.
technological options, policy design and strategy or to explore Barystatic and steric sea level changes do not include the effect of
the effects of additional impacts and adaptation beyond those changes in the shape of ocean basins induced by the change in the
represented in the reference scenario. ocean mass and its distribution. See also Vertical land motion (VLM),
Land water storage, Glacial isostatic adjustment (GIA), Extreme sea
Note 3: The term business as usual scenario has been used to describe
level (ESL) and Storm surge.
a scenario that assumes no additional policies beyond those currently
in place and that patterns of socio-economic development are Geocentric sea level change
consistent with recent trends. The term is now used less frequently The change in local mean sea surface height with respect to the
than in the past. terrestrial reference frame; it is the sea level change observed with
instruments from space. See also Altimetry.
Note 4: In climate change attribution or impact attribution research,
reference scenarios may refer to counterfactual historical scenarios AVII
251
Annex VII Glossary
Global mean sea level (GMSL) change Thermosteric sea level change
The increase or decrease in the volume of the ocean divided by Thermosteric sea level change (where thermosteric sea level rise may
the ocean surface area. It is the sum of changes in ocean density also be referred to as thermal expansion) occurs as a result of changes
through temperature changes (global mean thermosteric sea level in ocean temperature: increasing temperature reduces ocean density
change) and changes in the ocean mass as a result of changes in the and increases the volume per unit of mass.
cryosphere or land water storage (barystatic sea level change).
Sea level equivalent (SLE) The SLE of a mass of water, ice, or
Gravitational, rotational and deformational (GRD) effects water vapour is that mass, converted to a volume using a density of
Changes in Earth gravity, Earth rotation and viscoelastic solid Earth 1000 kg m–3, and divided by the present-day ocean surface area of
2
deformation (GRD) result from the redistribution of mass between 3.625 × 1000 m . Thus, 362.5 Gt of water mass added to the ocean
terrestrial ice and water reservoirs and the ocean. Contemporary correspond to 1 mm of global mean sea level rise.
terrestrial mass loss leads to elastic solid Earth uplift and a nearby
Sea level rise (SLR) See Sea level change (sea level rise/sea
relative sea level fall (for a single source of terrestrial mass loss this
level fall).
is within ~2000 km, for multiple sources the distance depends on
the interaction of the different relative sea level patterns). Farther Sea surface temperature (SST) The subsurface bulk
away (more than ~7000 km for a single source of terrestrial mass temperature in the top few metres of the ocean, measured by ships,
loss), relative sea level rises more than the global average, due (to buoys and drifters. From ships, measurements of water samples
first order) to gravitational effects. Earth deformation associated in buckets were mostly switched in the 1940s to samples from
with adding water to the oceans and a shift of the Earth’s rotation engine intake water. Satellite measurements of skin temperature
axis towards the source of terrestrial mass loss leads to second-order (uppermost layer; a fraction of a millimetre thick) in the infrared or
effects that increase spatial variability of the pattern globally. GRD the top centimetre or so in the microwave are also used, but must be
effects due to the redistribution of ocean water within the ocean adjusted to be compatible with the bulk temperature.
itself are referred to as self-attraction and loading effects.
Semi-direct (aerosol) effect See Aerosol–radiation interaction.
Halosteric sea level change
Semi-empirical model Model in which calculations are based
Halosteric sea level change occurs as a result of salinity variations:
on a combination of observed associations between variables and
higher salinity leads to higher density and decreases the volume per
theoretical considerations relating variables through fundamental
unit of mass. Although both processes can be relevant on regional to
principles (e.g., conservation of energy). For example, in sea level
local scales, only thermosteric changes impact the global mean sea
studies, semi-empirical models refer specifically to transfer functions
level (GMSL) change, whereas the global mean halosteric change is
formulated to project future global mean sea level (GMSL) change,
negligible (Gregory et al., 2019).
or contributions to it, from future global surface temperature change
Local sea level change or radiative forcing.
Change in sea level relative to a datum (such as present-day mean
Sensible heat flux The turbulent or conductive flux of heat from
sea level) at spatial scales smaller than 10 km.
the Earth’s surface to the atmosphere that is not associated with
Ocean dynamic sea level change phase changes of water; a component of the surface energy budget.
Change in mean sea level relative to the geoid associated with See also Latent heat flux.
circulation and density-driven changes in the ocean. Ocean dynamic
Sequestration The process of storing carbon in a carbon pool.
sea level change is regionally varying but by definition has a zero
See also Pool, carbon and nitrogen, Reservoir, Sequestration potential,
global mean and conventionally is inverse-barometer corrected (i.e., the
Sink, Source and Uptake.
effect of the hydrostatic depression of the sea surface by atmospheric
pressure changes is removed). Changes in ocean currents occur due to Sequestration potential The quantity of greenhouse gases that
variations in heating and cooling, variability in winds and changes in can be removed from the atmosphere by anthropogenic enhancement
seasonally to annually averaged air temperature and humidity. of sinks and stored in a pool. See Mitigation potential for different
subcategories of sequestration potential. See also Pool, carbon and
Regional sea level change
nitrogen, Reservoir, Sequestration, Source and Uptake.
Change in sea level relative to a datum (such as present-day mean
sea level) at spatial scales of about 100 km. Shared policy assumptions (SPAs) See Shared Socio-economic
Pathways (SSPs) (under Pathways).
Relative sea level (RSL) change
The change in local mean sea surface height (SSH) relative to the Shared Socio-economic Pathways (SSPs) See Pathways.
local solid surface, that is, the sea floor, as measured by instruments
that are fixed to the Earth’s surface, such as tide gauges. This Short-lived climate forcers (SLCFs) A set of chemically
reference frame is used when considering coastal impacts, hazards reactive compounds with short (relative to carbon dioxide (CO2))
and adaptation needs. atmospheric lifetimes (from hours to about two decades)
but characterized by different physiochemical properties and
Steric sea level change environmental effects. Their emission or formation has a significant
Steric sea level change is caused by changes in ocean density and effect on radiative forcing over a period determined by their
AVII is composed of thermosteric sea level change and halosteric sea respective atmospheric lifetimes. Changes in their emissions can also
level change. induce long-term climate effects via, in particular, their interactions
252
Glossary Annex VII
with some biogeochemical cycles. SLCFs are classified as direct or Snow cover extent (SCE)
indirect, with direct SLCFs exerting climate effects through their The areal extent of snow covered ground.
radiative forcing and indirect SLCFs being the precursors of other
Snow water equivalent (SWE)
direct climate forcers. Direct SLCFs include methane (CH4 ), ozone
The depth of liquid water that would result if a mass of snow melted
(O3 ), primary aerosols and some halogenated species. Indirect
completely.
SLCFs are precursors of ozone or secondary aerosols. SLCFs can be
cooling or warming through interactions with radiation and clouds. Socio-economic scenario See Scenario.
They are also referred to as near-term climate forcers. Many SLCFs
Soil moisture Water stored in the soil in liquid or frozen form.
are also air pollutants. A subset of exclusively warming SLCFs is
Root-zone soil moisture is of most relevance for plant activity.
also referred to as short-lived climate pollutants (SLCPs), including
methane, ozone, and black carbon (BC). Soil temperature The temperature of the soil. This can be
measured or modelled at multiple levels within the depth of the soil.
Short-lived climate pollutants (SLCP) See Short-lived climate
forcers (SLCFs). Solar activity General term collectively describing a variety of
magnetic phenomena on the Sun such as sunspots, faculae (bright
Shortwave radiation See Solar radiation.
areas), and flares (emission of high-energy particles). It varies on
Significant wave height The average trough-to-crest height of time scales from minutes to millions of years. The solar cycle, with an
the highest one-third of the wave heights (sea and swell) occurring in average duration of 11 years, is an example of a quasi-regular change
a particular time period. in solar activity.
Simple climate model (SCM) A broad class of lower- Solar cycle (11-year) A quasi-regular modulation of solar activity
dimensional models of the energy balance, radiative transfer, carbon with varying amplitude and a period of between 8 and 14 years.
cycle, or a combination of such physical components. SCMs are
Solar radiation Electromagnetic radiation emitted by the Sun
also suitable for performing emulations of climate-mean variables
with a spectrum close to that of a black body with a temperature of
of Earth system models (ESMs), given that their structural flexibility
5770 K. The radiation peaks in visible wavelengths. When compared
can capture both the parametric and structural uncertainties across
to the terrestrial radiation it is often referred to as shortwave
process-oriented ESM responses. They can also be used to test
radiation. See also Insolation and Total solar irradiance (TSI).
consistency across multiple lines of evidence with regard to climate
sensitivity ranges, transient climate responses (TCRs), transient Solar radiation modification (SRM) Refers to a range of
climate response to cumulative CO2 emissions (TCREs) and carbon radiation modification measures not related to greenhouse gas
cycle feedbacks. See also Emulators and Earth system model of (GHG) mitigation that seek to limit global warming. Most methods
intermediate complexity (EMIC). involve reducing the amount of incoming solar radiation reaching
the surface, but others also act on the longwave radiation budget by
Sink Any process, activity or mechanism which removes
reducing optical thickness and cloud lifetime.
a greenhouse gas, an aerosol or a precursor of a greenhouse gas
from the atmosphere (UNFCCC Article 1.8 (UNFCCC, 1992)). See also Cirrus cloud thinning (CCT)
Pool, carbon and nitrogen, Reservoir, Sequestration, Sequestration One of several radiation modification approaches to counter the
potential, Source and Uptake. warming caused by greenhouse gases (GHGs). In this approach, it
is proposed to reduce the amount of cirrus clouds by injecting ice
Small Island Developing States (SIDS) Small Island
nucleating substances in the upper troposphere. The reduction in
Developing States (SIDS), as recognized by the United Nations
cirrus clouds is expected to increase the amount of longwave cooling
OHRLLS (UN Office of the High Representative for the Least
to space resulting in a planetary cooling. Although cirrus cloud
Developed Countries, Landlocked Developing Countries and Small
thinning primarily affects the longwave radiation budget of our
Island Developing States), are a distinct group of developing countries
planet, it is often identified as one of the solar radiation modification
facing specific social, economic and environmental vulnerabilities
(SRM) approaches in the literature.
(UN-OHRLLS, 2011). They were recognized as a special case both for
their environment and development at the Rio Earth Summit in Brazil Marine cloud brightening (MCB)
in 1992. Fifty-eight countries and territories are presently classified One of several solar radiation modification (SRM) approaches to
as SIDS by the UN OHRLLS, with 38 being UN member states and increase the planetary albedo. In this approach, it is proposed to
20 being Non-UN Members or Associate Members of the Regional inject sea salt aerosols into persistent marine low clouds. This is
Commissions (UN-OHRLLS, 2018). expected to increase the cloud droplet concentration of these clouds
and their reflectivity.
Snow cover Snow cover refers to all the snow that has
accumulated on the ground at a given time (UNESCO/IASH/ Stratospheric aerosol injection (SAI)
WMO, 1970). One of several solar radiation modification (SRM) approaches to
increase the planetary albedo. In the approach, it is proposed to
Snow cover duration (SCD)
inject highly reflective aerosols such as sulphates into the lower
How long snow continuously remains on the land surface, or the
stratosphere. This is expected to increase the fraction of solar
period between snow-on and snow-off dates.
radiation deflected to space resulting in a planetary cooling. AVII
253
Annex VII Glossary
Solubility pump A physicochemical process that transports the term storyline is used both in connection to scenarios as related
dissolved inorganic carbon from the ocean’s surface to its interior. The to a future trajectory of the climate and human systems or to a
solubility pump is primarily driven by the solubility of carbon dioxide weather or climate event. In this context, storylines can be used
(CO2) (with more CO2 dissolving in colder water) and the large-scale, to describe plural, conditional possible futures or explanations
thermohaline patterns of ocean circulation. of a current situation, in contrast to single, definitive futures or
explanations.
Source Any process or activity which releases a greenhouse gas,
an aerosol or a precursor of a greenhouse gas into the atmosphere Physical climate storyline
(UNFCCC Article 1.9 (UNFCCC, 1992)). See also Pool, carbon and A self-consistent and plausible unfolding of a physical trajectory of
nitrogen, Reservoir, Sequestration, Sequestration potential, Sink the climate system, or a weather or climate event, on time scales
and Uptake. from hours to multiple decades (Shepherd et al., 2018). Through this,
storylines explore, illustrate and communicate uncertainties in the
South American monsoon (SAmerM) See Global monsoon.
climate system response to forcing and in internal variability.
South and Southeast Asian monsoon (SAsiaM) See
Scenario storyline
Global monsoon.
A narrative description of a scenario (or family of scenarios),
Southern Annular Mode (SAM) See Annular modes. highlighting the main scenario characteristics, relationships between
key driving forces and the dynamics of their evolution.
South Pacific Convergence Zone (SPCZ) A band of low-
level convergence, cloudiness and precipitation ranging from the Stratification Process of forming of layers of (ocean) water
west Pacific warm pool south-eastwards towards French Polynesia. with different properties such as salinity, density and temperature
It is one of the most significant features of subtropical Southern that act as barrier for water mixing. The strengthening of near-
Hemisphere climate. It shares some characteristics with the Inter- surface stratification generally results in warmer surface waters,
tropical Convergence Zone (ITCZ), but is more extratropical in nature, decreased oxygen levels in deeper water, and intensification of ocean
especially east of the International Date Line. acidification (OA) in the upper ocean.
Southern Oscillation See El Niño–Southern Oscillation (ENSO). Stratosphere The highly stratified region of the atmosphere
above the tropopause, extending to about 50 km altitude. See
Specific humidity The specific humidity specifies the ratio of
also Troposphere.
the mass of water vapour to the total mass of moist air. See also
Relative humidity. Stratospheric aerosol injection (SAI) See Solar radiation
modification (SRM).
Stadial or stade A brief period of regional climatic cooling during
a glacial or interglacial interval, often characterized by transient Stratosphere–troposphere exchange (STE) Stratosphere–
glacial advances. Stadials are generally of short duration (hundreds troposphere exchange (STE) is understood as the flux of air or trace
to a few thousand years) compared to glacial or interglacial constituents across the tropopause, including both directions: the
intervals (lasting many thousands to tens of thousands of years). stratosphere to troposphere transport (STT) and troposphere to
One example of a regional stadial event is based on millennial scale stratosphere transport (TST). STE is one of the key factors controlling
cooling recorded by oxygen isotope ratios in Greenland ice cores, the budgets of ozone, water vapour and other substances in both the
the so called “Greenland Stadials” (Johnsen et al., 1992). See also troposphere and the lower stratosphere.
Interstadial or interstade.
Stratospheric ozone Stratospheric ozone describes the ozone
Statistical downscaling See Downscaling. (O3 ) that resides in the stratosphere, the region of the atmosphere
which exists between 10 and 50 kilometres above the surface of the
Steric sea level change See Sea level change (sea level rise/
earth. Ninety percent of total-column ozone resides in the stratosphere.
sea level fall).
See also Ozone layer and Ozone-depleting substances (ODSs).
Storm surge The temporary increase, at a particular locality,
Stratospheric polar vortex A large-scale region of cold air
in the height of the sea due to extreme meteorological conditions
poleward of approximately 60 degrees that is contained by a strong
(low atmospheric pressure and/or strong winds). The storm surge is
westerly jet from the tropopause (8–10 km) to the stratopause
defined as being the excess above the level expected from the tidal
(50–60 km) and that forms in each hemisphere during the winter
variation alone at that time and place. See also Sea level change (sea
half-year. Planetary waves can temporarily disrupt the vortex,
level rise/sea level fall) and Extreme sea level (ESL).
producing easterly winds and rapid warming over polar regions in the
Storm tracks Originally, a term referring to the tracks of stratosphere, and leading to substantial weakening or breakdown of
individual cyclonic weather systems, but now often generalized the vortex.
to refer to the main regions where the tracks of extratropical
Stratospheric sounding unit (SSU) A three-channel infrared
disturbances occur as sequences of low (cyclonic) and high
sounder on operational U.S. National Oceanic and Atmospheric
(anticyclonic) pressure systems.
Administration (NOAA) polar-orbiting satellites. The three channels
Storyline A way of making sense of a situation or a series of are used to determine profiles of temperature in the stratosphere
AVII events through the construction of a set of explanatory elements. (AMS, 2021).
Usually, it is built on logical or causal reasoning. In climate research,
254
Glossary Annex VII
Streamflow Water flow within a river channel, for example, those derived from the temporal variation of the main modes of
expressed in m3 s–1. A synonym for river discharge. climate variability). They can also be obtained from principal
component analysis, singular value decomposition/maximum
Subduction Ocean process in which surface waters enter the
covariance analysis, clustering based on spatial recurrence
ocean interior from the surface mixed layer through Ekman pumping
criteria, etc. See also Section Atlas.3.1 of the AR6 WGI report and
and lateral advection. The latter occurs when surface waters are
Teleconnection.
advected to a region where the local surface layer is less dense and
therefore must slide below the surface layer, usually with no change Temperature overshoot Exceedance of a specified global
in density. warming level, followed by a decline to or below that level during
a specified period of time (e.g., before 2100). Sometimes the
Sudden stratospheric warming (SSW) A phenomena of rapid
magnitude and likelihood of the overshoot is also characterized.
warming in the stratosphere at high latitudes (sometimes more than
The overshoot duration can vary from one pathway to the next, but
50°C in 1–2 days) that can cause breakdown of stratospheric polar
in most overshoot pathways in the literature and as referred to as
vortices.
overshoot pathways in the AR6, the overshoot occurs over a period
Sulphur hexafluoride (SF6) SF6, a greenhouse gas (GHG), is of at least one decade and up to several decades. See also Pathways.
mainly used in heavy industry to insulate high-voltage equipment
Terrestrial radiation Radiation emitted by the Earth’s surface,
and to assist in the manufacturing of cable-cooling systems and
the atmosphere and clouds. It is also known as thermal infrared or
semiconductors.
longwave radiation and is to be distinguished from the near-infrared
Sunspots Dark areas on the Sun where strong magnetic fields radiation that is part of the solar spectrum. Infrared radiation, in
reduce the convection, causing a temperature reduction of about general, has a distinctive range of wavelengths (spectrum) longer
1500 K compared to the surrounding regions. The number of sunspots than the wavelength of the red light in the visible part of the spectrum.
is higher during periods of higher solar activity and varies in particular The spectrum of terrestrial radiation is almost entirely distinct from
with the solar cycle. that of shortwave or solar radiation because of the difference in
temperature between the Sun and the Earth–atmosphere system.
Surface air temperature See Land surface air temperature
(LSAT) and Global mean surface air temperature (GSAT). Thermal expansion See Steric sea level change (under Sea level
change (sea level rise/sea level fall)).
Surface mass balance (SMB) See Mass balance/budget (of
glaciers or ice sheets). Thermocline The layer of maximum vertical temperature gradient
in the ocean, lying between the surface ocean and the abyssal ocean.
Surface temperature See Global mean surface air temperature
In subtropical regions, its source waters are typically surface waters
(GSAT), Global mean surface temperature (GMST), Land surface air
at higher latitudes that have subducted (see Subduction) and moved
temperature (LSAT) and Sea surface temperature (SST).
equatorward. At high latitudes, it is sometimes absent, replaced by
Surprises A class of risk that can be defined as low-likelihood but a halocline, which is a layer of maximum vertical salinity gradient.
well-understood events and events that cannot be predicted with
Thermohaline circulation (THC) See Meridional overturning
current understanding (see Section 1.4.4.3 in AR6 WGI Chapter 1).
circulation (MOC).
Swash See Extreme sea level (ESL).
Thermokarst Process by which characteristic landforms result
Talik A layer or body of unfrozen ground in a permafrost area from thawing of ice-rich permafrost or melting of massive ice
due to a local anomaly in thermal, hydrological, hydrogeological or (IPA, 2005).
hydrochemical conditions (IPA, 2005).
Thermosteric See Sea level change (sea level rise/sea level fall).
Technical potential See Mitigation potential.
Tide gauge A device at a coastal or deep-sea location that
Teleconnection Association between climate variables at continuously measures the level of the sea with respect to the
widely separated, geographically fixed locations related to each adjacent land. Time averaging of the sea level so recorded gives the
other through physical processes and oceanic and/or atmospheric observed secular changes of the relative sea level.
dynamical pathways. Teleconnections can be caused by several
Time of emergence (ToE) Time when a specific anthropogenic
climate phenomena, such as Rossby wave-trains, mid-latitude jet and
signal related to climate change is statistically detected to emerge
storm track displacements, fluctuations of the Atlantic Meridional
from the background noise of natural climate variability in a reference
Overturning Circulation (AMOC), fluctuations of the Walker
period, for a specific region (Hawkins and Sutton, 2012). See also
circulation, etc. They can be initiated by modes of climate variability,
Emergence (of the climate signal).
thus providing the development of remote climate anomalies at
various temporal lags. See also Teleconnection pattern. Tipping element A component of the Earth system that is
susceptible to a tipping point.
Teleconnection pattern Spatial structure of climate anomalies
that are linked to each other through teleconnection processes or Tipping point A critical threshold beyond which a system
that are the large-scale fingerprint of modes of climate variability. reorganizes, often abruptly and/or irreversibly. See also Tipping
Teleconnection patterns can be visualized using correlation and/or element, Irreversibility and Abrupt change. AVII
regression maps of climate variables with some climate indices (i.e.,
255
Annex VII Glossary
Total alkalinity Total Alkalinity (AT) is a measurable parameter of Atlantic Meridional Mode (AMM) The Atlantic Meridional Mode
the seawater acid–base system which, when expressed in micromoles (AMM) refers to the interannual to decadal variability of the cross-
per kilogram of seawater, is a conservative variable both on mixing equatorial sea surface temperature gradients and surface wind
and for changes in temperature and/or pressure. Changes in total anomalies in the tropical Atlantic. It modulates the strength and
alkalinity in the oceans can result from a variety of biogeochemical latitudinal shifts of the Inter-tropical Convergence Zone (ITCZ),
processes that affect the acid–base composition of the seawater which impacts regional rainfall over Northeast Brazil and Atlantic
itself. However, its value is not affected by the exchange of carbon hurricane activity. See Section AIV.2.5 in Annex IV of the AR6
dioxide gas between seawater and the atmosphere. Measurements of WGI report.
total alkalinity can thus be used to help study these biogeochemical
Atlantic Zonal Mode (AZM) An equatorial coupled mode in the
processes and can also be used to help calculate the state of the
Atlantic similar to El Niño–Southern Oscillation (ENSO) in the Pacific,
seawater acid–base system. Total alkalinity is most commonly
and therefore sometimes referred to as the Atlantic Niño. The AZM
measured using an acidimetric titration technique that determines
is associated with sea surface temperature anomalies near the
how much acid is required to titrate a seawater sample to a specified
equatorial Atlantic and rainfall disturbances over the African monsoon
equivalence point.
domain. Its variations are mostly observed in the interannual scale. It
Total carbon budget See Carbon budget. is called also Atlantic equatorial mode. See Section AIV.2.5 in Annex
IV of the AR6 WGI report.
Total solar irradiance (TSI) The total amount of solar radiation
in watts per square metre received outside the Earth’s atmosphere Tropical cyclone The general term for a strong, cyclonic-scale
on a surface normal to the incident radiation and at the Earth’s mean disturbance that originates over tropical oceans. Distinguished from
distance from the Sun. Reliable measurements of solar radiation can weaker systems (often named tropical disturbances or depressions)
only be made from space, and the precise record extends back only by exceeding a threshold wind speed. A tropical storm is a tropical
to 1978. Variations of a few tenths of a percent are common, usually cyclone with one-minute average surface winds between 18 and
associated with the passage of sunspots across the solar disk. The 32 m s–1. Beyond 32 m s–1, a tropical cyclone is called a hurricane,
solar cycle variation of TSI is of the order of 0.1% (AMS, 2021). See typhoon, or cyclone, depending on geographic location.
also Insolation.
Tropopause The boundary between the troposphere and the
Total water level See Extreme sea level (ESL). stratosphere. It ranges from 8–9 km at high latitudes to 15–16 km
in the tropics.
Trace gas A minor constituent of the atmosphere, next to
nitrogen and oxygen that together make up 99% of all volume. Troposphere The lowest part of the atmosphere, below the
The most important trace gases contributing to the greenhouse tropopause, where clouds and weather phenomena occur. In the
effect are carbon dioxide (CO2), ozone (O3 ), methane (CH4 ), nitrous troposphere, temperatures generally decrease with height. See
oxide (N2O), perfluorocarbons (PFCs), chlorofluorocarbons (CFCs), also Stratosphere.
hydrofluorocarbons (HFCs), sulphur hexafluoride (SF6 ) and water
Tropospheric ozone See Ozone (O3 ) and Ground-level ozone.
vapour (H2O).
Tundra A treeless biome characteristic of polar and alpine regions.
Transient climate response (TCR) See Climate sensitivity.
Turnover time (T) See Lifetime.
Transient climate response to cumulative CO2 emissions
(TCRE) See Climate sensitivity. Typhoon See Tropical cyclone.
Tree rings Concentric rings of secondary wood evident in a cross Typological domains See Typological regions.
section of the stem of a woody plant. The difference between the
Typological regions Regions of the Earth that share one or more
dense, small-celled late wood of one season and the wide-celled
specific features (known as ‘typologies’), such as geographic location
early wood of the following spring enables the age of a tree to be
(e.g., coastal), physical processes (e.g., monsoons), and biological
estimated, and the ring widths or density can be related to climate
(e.g., coral reefs, tropical forests), geological (e.g., mountains) or
parameters such as temperature and precipitation.
anthropogenic (e.g., megacities) formation, and for which it is useful
Tropical Atlantic modes See Tropical Atlantic Variability (TAV). to consider the common climate features. Typological regions are
smaller than climatic zones (e.g., a mountain region) and can be
Tropical Atlantic Variability (TAV) A generic term to describe
discontinuous (e.g., a group of megacities affected by the urban heat
the climate variability of the tropical Atlantic which is dominated
island effect, or monsoon regions).
at interannual to decadal time scales by two main climate modes:
the Atlantic Zonal Mode (AZM) and the Atlantic Meridional Mode Uncertainty A state of incomplete knowledge that can result
(AMM). The Atlantic Zonal Mode, also commonly referred to as the from a lack of information or from disagreement about what is
Atlantic Niño or Atlantic equatorial mode, is associated with sea known or even knowable. It may have many types of sources,
surface temperature anomalies near the equator, peaking in the from imprecision in the data to ambiguously defined concepts
eastern basin, while the Atlantic meridional mode is characterized or terminology, incomplete understanding of critical processes, or
by an inter-hemispheric gradient of sea surface temperature uncertain projections of human behaviour. Uncertainty can therefore
AVII and wind anomalies. Both modes are associated with significant be represented by quantitative measures (e.g., a probability density
teleconnections over Africa and South America. function) or by qualitative statements (e.g., reflecting the judgement
256
Glossary Annex VII
of a team of experts) (see Moss and Schneider, 2000; IPCC, 2004; Vertical land motion (VLM) The change in height of the land
Mastrandrea et al., 2010). See also Confidence and Likelihood. surface or the sea floor and can have several causes in addition
to elastic deformation associated with contemporary changes in
Deep uncertainty
gravity, rotation and viscoelastic solid Earth deformation (GRD) and
A situation of deep uncertainty exists when experts or stakeholders
viscoelastic deformation associated with glacial isostatic adjustment
do not know or cannot agree on: (1) appropriate conceptual models
(GIA). Subsidence (sinking of the land surface or sea floor) can, for
that describe relationships among key driving forces in a system;
instance, occur through compaction of alluvial sediments in deltaic
(2) the probability distributions used to represent uncertainty about
regions, removal of fluids such as gas, oil, and water, or drainage
key variables and parameters; and/or (3) how to weigh and value
of peatlands. Tectonic deformation of the Earth’s crust can occur as
desirable alternative outcomes (Lempert et al., 2003).
a result of earthquakes and volcanic eruptions. See also Sea level
Interpolation uncertainty change (sea level rise/sea level fall).
Uncertainty arising from a statistical or physical model-based
Very short-lived halogenated substances (VSLSs) Very
interpolation of a field between available estimates to create a more
short-lived halogenated substances (VSLSs) are considered to include
spatio-temporally complete estimate.
source gases (very short-lived halogenated substances present in
Sampling uncertainty the atmosphere in the form they were emitted from natural and
Uncertainty arising from incomplete or uneven availability of anthropogenic sources), halogenated product gases arising from
measurements in either space or time or both. source gas degradation, and other sources of tropospheric inorganic
halogens. VSLSs have tropospheric lifetimes of around 0.5 years
Trend estimates uncertainty
or less.
Uncertainty arising from data fitting to a time-series with potential
non-linear and autorogressive character. Volatile organic compounds (VOCs) Important class of organic
chemical air pollutants that are volatile at ambient air conditions.
United Nations Framework Convention on Climate Change
Other terms used to represent VOCs are hydrocarbons (HCs), reactive
(UNFCCC) The UNFCCC was adopted in May 1992 and opened
organic gases (ROGs) and non-methane volatile organic compounds
for signature at the 1992 Earth Summit in Rio de Janeiro. It entered
(NMVOCs). NMVOCs are major contributors – together with nitrogen
into force in March 1994 and as of September 2020 had 197 Parties
oxides (NOx), and carbon monoxide (CO) – to the formation of
(196 States and the European Union). The Convention’s ultimate
photochemical oxidants such as ozone (O3 ).
objective is the ‘stabilization of greenhouse gas concentrations in the
atmosphere at a level that would prevent dangerous anthropogenic Biogenic volatile organic compounds (BVOCs)
interference with the climate system’ (UNFCCC, 1992). The provisions Organic gas-phase compounds emitted from terrestrial and aquatic
of the Convention are pursued and implemented by two further ecosystems that are critical in ecology and plant physiology, from
treaties: the Kyoto Protocol and the Paris Agreement. abiotic and biotic stress functions to integrated components of
metabolism. BVOCs are important in atmospheric chemistry as
Uptake The transfer of substances (such as carbon) or energy (e.g.,
precursors for ozone (O3 ) and secondary organic aerosol formation.
heat) from one compartment of a system to another; for example,
Other terms used to represent BVOCs are hydrocarbons (HCs),
in the Earth system from the atmosphere to the ocean or to the
reactive organic gases (ROGs) and non-methane volatile organic
land. See also Pool, carbon and nitrogen, Reservoir, Sequestration,
compounds (NMVOCs).
Sequestration potential, Sink and Source.
Vulnerability The propensity or predisposition to be adversely
Upwelling region A region of an ocean where cold, typically
affected. Vulnerability encompasses a variety of concepts and
nutrient-rich waters well up from the deep ocean.
elements including sensitivity or susceptibility to harm and lack of
Urban heat island (UHI) The relative warmth of a city compared capacity to cope and adapt. See also Exposure, Hazard and Risk.
with surrounding rural areas, associated with heat trapping due to
Walker circulation Direct thermally driven zonal overturning
the close proximity of tall buildings, the heat-absorbing properties
circulation in the atmosphere over the tropical Pacific Ocean, with
of urban building materials, reduced ventilation, and heat generated
rising air in the western and sinking air in the eastern Pacific.
directly from human activities. See also Urbanization.
Warm spell See Heatwave.
Urbanization In the WGI report, urbanization is used to mean the
process of soil sealing with the change of natural land cover to built Water cycle See Hydrological cycle.
environment and urban areas, together with its associated albedo
Water mass A body of ocean water with identifiable properties
changes, and increased surface runoff and elevated warming. See
(temperature, salinity, density, chemical tracers) resulting from its
also Urban heat island (UHI).
unique formation process. Water masses are often identified through
Ventilation The exchange of ocean properties with the a vertical or horizontal extremum of a property such as salinity. North
atmospheric surface layer such that property concentrations are Pacific Intermediate Water (NPIW) and Antarctic Intermediate Water
brought closer to equilibrium values with the atmosphere (AMS, (AAIW) are examples of water masses.
2021), and the processes that propagate these properties into the
Water security ‘The capacity of a population to safeguard
ocean interior.
sustainable access to adequate quantities of acceptable quality AVII
water for sustaining livelihoods, human well-being, and socio-
257
Annex VII Glossary
AVII
258
Glossary Annex VII
259
Annex VII Glossary
Mastrandrea, M.D. et al., 2010: Guidance Note for Lead Authors of the land-use-land-use-change-and-forestry-lulucf/reporting-and-accounting-
IPCC Fifth Assessment Report on Consistent Treatment of Uncertainties. of-lulucf-activities-under-the-kyoto-protocol.
Intergovernmental Panel on Climate Change (IPCC), Geneva, Switzerland, UNFCCC, 2021b: Reporting and Review under the Paris Agreement. United
6 pp., www.ipcc.ch/publication/ipcc-cross-working-group-meeting-on- Nations Framework Convention on Climate Change (UNFCCC). Retrieved
consistent-treatment-of-uncertainties. from: https://unfccc.int/process-and-meetings/transparency-and-reporting/
Mix, A.C., W.F. Ruddiman, and A. McIntyre, 1986: Late Quaternary reporting-and-review-under-the-paris-agreement.
paleoceanography of the Tropical Atlantic, 1: Spatial variability of annual UNGA, 2016: Report of the open-ended intergovernmental expert working
mean sea-surface temperatures, 0–20,000 years B.P.. Paleoceanography, group on indicators and terminology relating to disaster risk reduction.
1(1), 43–66, doi:10.1029/pa001i001p00043. A/71/644, United Nations General Assembly (UNGA), 41 pp., https://
Mix, A.C., E. Bard, and R. Schneider, 2001: Environmental processes of the ice digitallibrary.un.org/record/852089.
age: land, oceans, glaciers (EPILOG). Quaternary Science Reviews, 20(4), UN-OHRLLS, 2011: Small Island Developing States: Small Islands Big(ger)
627–657, doi:10.1016/s0277-3791(00)00145-1. Stakes. Office for the High Representative for the Least Developed
Moss, R.H. and S.H. Schneider, 2000: Uncertainties in the IPCC TAR: Countries, Landlocked Developing Countries and Small Island Developing
Recommendations to Lead Authors for More Consistent Assessment and States (UN-OHRLLS), New York, NY, USA, 32 pp.
Reporting. In: Guidance Papers on the Cross Cutting Issues of the Third UN-OHRLLS, 2018: Small Island Developing States: Country profiles. Office
Assessment Report of the IPCC [Pachauri, R., T. Taniguchi, and K. Tanaka for the High Representative for the Least Developed Countries, Landlocked
(eds.)]. Intergovernmental Panel on Climate Change (IPCC), Geneva, Developing Countries and Small Island Developing States (UN-OHRLLS).
Switzerland, pp. 33–51. Retrieved from: http://unohrlls.org/about-sids/country-profiles.
Moss, R.H. et al., 2010: The next generation of scenarios for climate UN-Water, 2013: What is Water Security? Infographic. UN-Water, Geneva,
change research and assessment. Nature, 463(7282), 747–756, Switzerland. Retrieved from: www.unwater.org/publications/water-
doi:10.1038/nature08823. security-infographic.
NOAA, 2021: What is an iceberg? National Oceanic and Atmospheric Walker, M. et al., 2019: Formal Subdivision of the Holocene Series/Epoch:
Administration (NOAA). National Ocean Service website. Retrieved from: A Summary. Journal of the Geological Society of India, 93(2), 135–141,
https://oceanservice.noaa.gov/facts/iceberg.html. doi:10.1007/s12594-019-1141-9.
Pongratz, J. et al., 2018: Models meet data: Challenges and opportunities in Wiener, J.B. and J.D. Graham (eds.), 2009: Risk vs Risk: Tradeoffs in Protecting
implementing land management in Earth system models. Global Change Health and the Environment. Harvard University Press, Cambridge, MA,
Biology, 24(4), 1470–1487, doi:10.1111/gcb.13988. USA, 352 pp.
Ralph, F.M., M.D. Dettinger, M.M. Cairns, T.J. Galarneau, and J. Eylander, 2018: Yokoyama, Y. et al., 2018: Rapid glaciation and a two-step sea level
Defining “Atmospheric River”: How the Glossary of Meteorology Helped plunge into the Last Glacial Maximum. Nature, 559(7715), 603–607,
Resolve a Debate. Bulletin of the American Meteorological Society, 99(4), doi:10.1038/s41586-018-0335-4.
837–839, doi:10.1175/bams-d-17-0157.1. Zscheischler, J. et al., 2018: Future climate risk from compound events. Nature
Schwartz, S.E. and P. Warneck, 1995: Units for use in atmospheric chemistry Climate Change, 8(6), 469–477, doi:10.1038/s41558-018-0156-3.
(IUPAC Recommendations 1995). Pure and Applied Chemistry, 67(8/9),
1377–1406, http://publications.iupac.org/pac/1995/pdf/6708x1377.pdf.
Shepherd, T.G. et al., 2018: Storylines: an alternative approach to representing
uncertainty in physical aspects of climate change. Climatic Change,
151(3–4), 555–571, doi:10.1007/s10584-018-2317-9.
Steffen, W. et al., 2016: Stratigraphic and Earth System approaches
to defining the Anthropocene. Earth’s Future, 4(8), 324–345,
doi:10.1002/2016ef000379.
Stocker, T.F. and S.J. Johnsen, 2003: A minimum thermodynamic
model for the bipolar seesaw. Paleoceanography, 18(4), 1087,
doi:10.1029/2003pa000920.
Türkeş, M., 1999: Vulnerability of Turkey to Desertification With Respect to
Precipitation and Aridity Conditions. Turkish Journal of Engineering and
Environmental Sciences, 23, 363–380.
UN, 1992: Article 2: Use of Terms. In: Convention on Biological Diversity.
United Nations (UN), pp. 3–4, www.cbd.int/doc/legal/cbd-en.pdf.
UNCCD, 1994: United Nations Convention to Combat Desertification in
countries experiencing serious drought and/or desertification, particularly in
Africa. 58 pp., https://treaties.un.org/doc/Treaties/1996/12/19961226%20
01-46%20PM/Ch_XXVII_10p.pdf.
UNESCO/IASH/WMO, 1970: Seasonal snow cover: A guide for measurement,
compilation and assemblage of data. United Nations Educational, Scientific
and Cultural Organization (UNESCO), Paris, France, 38 pp.
UNFCCC, 1992: United Nations Framework Convention on Climate Change.
FCCC/INFORMAL/84, United Nations Framework Convention on Climate
Change (UNFCCC), 24 pp., https://unfccc.int/resource/docs/convkp/
conveng.pdf.
UNFCCC, 2021a: Reporting and accounting of LULUCF activities under the
Kyoto Protocol. United Nations Framework Convention on Climate Change
AVII (UNFCCC). Retrieved from: https://unfccc.int/topics/land-use/workstreams/
260