Download as pdf or txt
Download as pdf or txt
You are on page 1of 148

Institute of Rural Management Anand

Sustainable Development and Green Business


(SDGB)

PGDM (RM)- 41(2020–2022)

Term III
Course Facilitator:

Pramod K. Singh

January 2020
Email: pramod@irma.ac.in Mob. 9662514668
ii
PGDM (RM)- 41(2020–2022)
Academic Year (2020-21)

Course Title: Sustainable Development and Green Business (SDGB)


Course Instructor Pramod K. Singh
Term III
Credit 1.5 (15 hours)
Core / Elective Elective
Introduction The course discusses the problems of sustainable development with trans-
disciplinary perspectives. It deals with the interactions between natural and
social systems, and with how those interactions affect the challenge of
sustainability: meeting the needs of present and future generations while
substantially reducing poverty and conserving the planet’s life support systems.
The course presents a perspective as to how economic development and
human well-being can be decoupled from the destruction of natural resources.

Corporations print annual “Sustainability Reports,” insert the term into press
releases and CEO speeches, create new positions such as the Chief
Sustainability Officer, and gather for conferences on the “sustainability
challenge.” A survey by Price Waterhouse Coopers found that 87% of Fortune
1000 CEOs believe sustainability is important to a company’s profits.

Over time, sustainability has become less and less an isolated business
concern. The firm's business channels have been altered to bring
environmental and social issues to managerial attention through avenues
related to marketing, accounting, finance, product development, etc. For each
case, firms have pre-existing models and language that enable them to
understand the issue and formulate a response. As these responses have
become routinized, ongoing sustainability issues are treated as common
strategic concerns, no longer dictated by external social interests or ecosystem
constraints, but rather by internal strategic norms.

The course is inter-disciplinary in nature. It draws from the fields of natural and
social sciences such as ecology and environmental sciences, agriculture,
climate science and policy, institutional economics, community development,
gender studies, etc. Module three integrates these concepts for attaining long-
term human well-being. The course tries to analyse how a business could
promote sustainable development.

Course Objectives a. To build perspective on interactions between natural and social systems,
affecting the challenge of sustainable development.
b. To develop a deeper understanding of building synergies between SDGs
c. To develop a deeper understanding of climate change mitigation,
adaptation, and governance.
d. To impart managerial know-how for building climate-smart primary
production systems.
e. To understand and analyse the changing environmental strategies of
corporations across the globe
f. To evaluate claims of cleaner production and green business strategies by
the industry.
g. To develop perspectives for attaining long-term human well-being

iii
Connect with The course helps the participants develop perspectives on IRMA’s mission of
IRMA’s Mission sustainable, ecologically friendly, and equitable socio-economic development.
and Course India is committed to achieving Sustainable development goals (SDGs). The
course addresses ten out of seventeen SDGs. These SDGs are: 6, 7, 9, 11, 12,
13, 14, 15, & 17.
Learning • Building perspective on interactions between natural and social systems,
Outcomes affecting the challenge of sustainable development.
• Developing a deeper understanding of building synergies between SDGs
• Developing a deeper understanding of climate change mitigation,
adaptation, and governance.
• Imparting managerial know-how for building climate-smart primary
production systems.
• Understanding and analysing the changing environmental strategies of
corporations across the globe
• Evaluating claims of cleaner production and green business strategies
by the industry.
• Developing perspectives for attaining long-term human well-being
Measuring
Learning Components Weights (%)
Outcomes & Group Assignment 20
Evaluation Plan Individual Assignment 10
Group Project 20
End-Term Examination (Open book) 40
Class Participation 10
Total 100
Session wise Plan Session Plan
with Readings / Sessi Topics Readings Cases
References on
Numb
er
Module 1: Attaining Sustainable Development: Synergies and Trade-
Offs (4 sessions)
1 Concepts of sustainable development; R1, R2
Diagnosis of the unsustainable trends of the
contemporary world; Building synergies
between SDGs
2 The philosophical basis for sustainable R3
development; Green Development Concepts
and Environmental Movements (Deep
Ecology, Gaia hypothesis, Eco-feminism, etc.)
3-4 Strategies for human survival in the context of R4, R5
climate change: adaptation and mitigation
Module 2: Green Business: The Changing Business Landscape (4
sessions)
5 Understanding and analysing green and R6, R7
responsible business throughout the product
life cycle
6 Triple Bottom Line strategy for green and R6, R7
responsible business
6 Deceiving corporate strategy - Greenwashing R6, R7

iv
7 Emerging strategies of Dematerialisation and R8
Decarbonisation
8 Green marketing strategy R6, R7
9 Industrial system transition for a R9
dematerialised and decarbonised world
Module 3: Long-term sustainable development trajectories (2 sessions)
10 Climate resilient development: human well- R10
being within a safe and just space in the 21st
Century
Terminal Examination
Simulation/Games Instructor’s designed exercises
/ Exercises

Compulsory Readings
R Session Reading Title Pages

Module 1: Attaining Sustainable Development: Synergies and Trade-Offs

R1 1 Definitions of important terminologies (compiled by the IPCC in 2020; 1-33


based on the contributions of author teams)
Note: This reading is essential for all the sessions
R2 1 United Nations (2015) Transforming our world: the 2030 Agenda for 34-49
Sustainable Development
R3 4 Environmental Movements (Compiled from various sources by the 50-56
instructor for class discussion only)
R4 2 IPCC (2018) Global Warming of 1.5o C: Technical Summary 57-72
R5 2 Understanding Climate Change Mitigation, Adaptation and 73-78
Governance (A note by the instructor)
Module 2: Green Business: The Changing Business Landscape
R6 5 Key Green Business Concepts (Compiled from various sources by the 79-85
instructor for class discussion only)
Note: This reading is essential for sessions 5-9
R7 5 Andrew J. Hoffman (2018) The next phase of business sustainability, 86-92
Stanford Social Innovation Review, Spring 2018.

R8 7 Global Environment Outlook 2019, Chapter 17 Towards the Circular 93-97


Economy (part of the chapter)
R9 9 Industrial system transition for a dematerialised and decarbonised 99-114
world (a manuscript by Singh and Chudasama)
Module 3: Long-term sustainable development trajectory
R10 10 Climate resilient development: human well-being within a safe and 115-142
just space in the 21st Century (a manuscript by Singh and
Chudasama)

v
Cases

C Session Case Title


C1 5 Triple Bottom-line strategy (Amanco: Developing Sustainability
Scorecard)
C2 5 Deceiving corporate strategy - Greenwashing (Fiji Water)
C3 6 Life Cycle Approach (IBM)
C4 6 Dematerialisation strategy (Dow)
C5 7 Decarbonisation strategy (Elon Musk’s Big Bets)
C6 7 Decarbonisation strategy in emerging technology (New Light
Technologies: Plastics for a Carbon Negative Future)
C7 8 Market transformation strategy (Wall Mart)

Note: Cases will be supplied electronically after session 3

vi
01 Definitions of Important Terminologies
(Internal) Displacement (of humans): The involuntary movement, individually or collectively, of persons from
their country or community, notably for reasons of armed conflict, civil unrest, or natural or man-made disasters
(adapted from IOM, 2011).
2030 Agenda for Sustainable Development: A UN resolution in September 2015 adopting a plan of action for
people, planet and prosperity in a new global development framework anchored in 17 Sustainable Development
Goals (UN, 2015).
Abrupt change / abrupt climate change: Abrupt change refers to a change that is substantially faster than the rate
of change in the recent history of the affected components of a system. Abrupt climate change refers to a regional to
global scale change in the climate system that occurs faster than the rate of change of forcing, implying non-
linearity in the climate response. This definition includes shifts from one equilibrium state to another (tipping
points), but also other non-linear responses of the climate system. Abrupt climate changes take place over a few
decades or less, persist (or are anticipated to persist) for at least a few decades, and cause substantial disruptions in
human and natural systems. They are sometimes called rapid climate changes, abrupt events or even surprises.
Access to food: See Access under Food Security
Acclimatization: A change in functional or morphological traits occurring once or repeatedly (e.g., seasonally)
during the lifetime of an individual organism in its natural environment. Through acclimatization, the individual
maintains performance across a range of environmental conditions. For a clear differentiation between findings in
laboratory and field studies, the term acclimation is used in ecophysiology for the respective phenomena when
observed in well-defined experimental settings. The term (adaptive) plasticity characterises the generally limited
scope of changes in phenotype that an individual can reach through the process of acclimatization.
Accumulation (of glaciers, ice sheets, or snow cover): All processes that add to the mass of a glacier, an ice
sheet, or snow cover. The main process of accumulation is snowfall. Accumulation also includes deposition of
hoar, freezing rain, other types of solid precipitation, gain of wind-blown snow, avalanching, and basal
accumulation (often beneath floating ice).
Adaptation: In human systems, the process of adjustment to actual or expected climate and its effects, in order to
moderate harm or exploit beneficial opportunities. In natural systems, the process of adjustment to actual climate
and its effects; human intervention may facilitate adjustment to expected climate and its effects.
Adaptation assessment: The practice of identifying options to adapt to climate change and evaluating them in
terms of criteria such as availability, benefits, costs, effectiveness, efficiency, and feasibility.
Adaptation deficit: The gap between the current state of a system and a state that minimizes adverse impacts from
existing climate conditions and variability.
Adaptation gap: The difference between actually implemented adaptation and a societally set goal, determined
largely by preferences related to tolerated climate change impacts and reflecting resource limitations and competing
priorities” (UNEP, 2014; UNEP, 2018).
Adaptation limits: The point at which an actor’s objectives (or system needs) cannot be secured from intolerable
risks through adaptive actions.
•Hard adaptation limit - No adaptive actions are possible to avoid intolerable risks.
•Soft adaptation limit - Options may exist but are currently not available to avoid intolerable
risks through adaptive action.
Adaptation needs: The circumstances requiring action to ensure safety of populations and security of assets in
response to climate impacts.
Adaptation options: The array of strategies and measures that are available and appropriate for addressing
adaptation. They include a wide range of actions that can be categorized as structural, institutional, ecological or
behavioural.

1
Autonomous adaptation: Adaptation in response to experienced climate and its effects, without planning
explicitly or consciously focused on addressing climate change. Also referred to as spontaneous adaptation.
Community-based adaptation: Local, community-driven adaptation. Community-based adaptation focuses
attention on empowering and promoting the adaptive capacity of communities. It is an approach that takes context,
culture, knowledge, agency, and preferences of communities as strengths.
Ecosystem-based adaptation (EBA): The use of ecosystem management activities to increase the resilience and
reduce the vulnerability of people and ecosystems to climate change (Campbell et al., 2009).
See also Nature-based solutions
Evolutionary adaptation: The process whereby a species or population becomes better able to live in a changing
environment, through the selection of heritable traits. Biologists usually distinguish evolutionary adaptation from
acclimatization, with the latter occurring within an organism’s lifetime.
Incremental adaptation: Adaptation that maintains the essence and integrity of a system or process at a given
scale (Park et al., 2012). In some cases, incremental adaptation can accrue to result in transformational adaptation
(Tàbara et al., 2018; Termeer et al., 2017). Incremental adaptations to change in climate are understood as
extensions of actions and behaviours that already reduce the losses or enhance the benefits of natural variations in
extreme weather / climate events.
Transformational adaptation: Adaptation that changes the fundamental attributes of a social-ecological system in
anticipation of climate change and its impacts.
Adaptation Fund: A Fund established under the Kyoto Protocol in 2001 and officially launched in 2007. The Fund
finances adaptation projects and programmes in developing countries that are Parties to the Kyoto protocol.
Financing comes mainly from sales of Certified Emissions Reductions (CERs) and a share of proceeds amounting
to 2% of the value of CERs issued each year for Clean Development Mechanism (CDM) projects. The Adaptation
Fund can also receive funds from government, private sector, and individuals.
Adaptation limits: See Adaptation
Adaptation needs: See Adaptation
Adaptation options: See Adaptation
Adaptation pathways: See Pathways
Adaptive capacity: The ability of systems, institutions, humans and other organisms to adjust to potential damage,
to take advantage of opportunities, or to respond to consequences (IPCC, 2014; MA, 2005).
Adaptive governance: See Governance
Adaptive management: A process of iteratively planning, implementing, and modifying strategies for managing
resources in the face of uncertainty and change. Adaptive management involves adjusting approaches in response to
observations of their effect and changes in the system brought on by resulting feedback effects and other variables.
Aerosol: A suspension of airborne solid or liquid particles, with a typical size between a few nanometres and 10
μm, that reside in the atmosphere for at least several hours. The term aerosol, which includes both the particles and
the suspending gas, is often used in this report in its plural form to mean ‘aerosol particles’. Aerosols may be of
either natural or anthropogenic origin. Aerosols can influence climate in several ways: directly through scattering
and absorbing radiation, and indirectly by acting as cloud condensation nuclei or ice nuclei, modifying the optical
properties and lifetime of clouds or upon deposition on snow- or ice-covered surfaces thereby altering their albedo
and contributing to climate feedback. Atmospheric aerosols, whether natural or anthropogenic, originate from two
different pathways: emissions of primary particulate matter (PM), and formation of secondary PM from gaseous
precursors. The bulk of aerosols are of natural origin. Some scientists use group labels that refer to the chemical
composition, namely: sea salt, organic carbon, black carbon (BC), mineral species (mainly desert dust), sulphate,
nitrate, and ammonium. These labels are, however, imperfect as aerosols combine particles to create complex
mixtures. See also Short-lived climate forcers (SLCF).
Afforestation: Conversion to forest of land that historically has not contained forests.

2
[Note: For a discussion of the term forest and related terms such as afforestation, reforestation and deforestation, in
the context of reporting and accounting Article 3.3 and 3.4 activities under the Kyoto Protocol, see 2013 Revised
Supplementary Methods and Good Practice Guidance Arising from the Kyoto Protocol.]
Agreement: In this report, the degree of agreement within the scientific body of knowledge on a particular finding
is assessed based on multiple lines of evidence (e.g., mechanistic understanding, theory, data, models, expert
judgement) and expressed qualitatively (Mastrandrea et al., 2010). See also Confidence, Likelihood and
Uncertainty.
Agroecology: The science and practice of applying ecological concepts, principles and knowledge (i.e., the
interactions of, and explanations for, the diversity, abundance and activities of organisms) to the study, design and
management of sustainable agroecosystems. It includes the roles of human beings as a central organism in
agroecology by way of social and economic processes in farming systems. Agroecology examines the roles and
interactions among all relevant biophysical, technical and socioeconomic components of farming systems and their
surrounding landscapes (IPBES, 2019).
Agroforestry: Collective name for land-use systems and technologies where woody perennials (trees, shrubs,
palms, bamboos, etc.) are deliberately used on the same land-management units as agricultural crops and/or
animals, in some form of spatial arrangement or temporal sequence. In agroforestry systems there are both
ecological and economical interactions between the different components. Agroforestry can also be defined as a
dynamic, ecologically based, natural resource management system that, through the integration of trees on farms
and in the agricultural landscape, diversifies and sustains production for increased social, economic and
environmental benefits for land users at all levels (FAO, 2015a).
Air pollution: Degradation of air quality with negative effects on human health or the natural or built environment
due to the introduction, by natural processes or human activity, into the atmosphere of substances (gases, aerosols)
which have a direct (primary pollutants) or indirect (secondary pollutants) harmful effect. See also Short- lived
climate forcers (SLCF).
Albedo: The proportion of sunlight (solar radiation) reflected by a surface or object, often expressed as a
percentage.
Clouds, snow and ice usually have high albedo; soil surfaces cover the albedo range from high to low; vegetation in
the dry season and/or in arid zones can have high albedo, whereas photosynthetically active vegetation and the
ocean have low albedo. The Earth's planetary albedo changes mainly through changes in cloudiness, snow, ice, leaf
area and land cover.
Anomaly: The deviation of a variable from its value averaged over a reference period.
Anthropocene: A proposed new geological epoch resulting from significant human-driven changes to the structure
and functioning of the Earth System, including the climate system. Originally proposed in the Earth System science
community in 2000, the proposed new epoch is undergoing a formalization process within the geological ommunity
based on the stratigraphic evidence that human activities have changed the Earth System to the extent of forming
geological deposits with a signature that is distinct from those of the Holocene, and which will remain in the
geological record. Both the stratigraphic and Earth System approaches to defining the Anthropocene consider the
mid-20th Century to be the most appropriate starting date, although others have been proposed and continue to be
discussed. The Anthropocene concept has been taken up by a diversity of disciplines and the public to denote the
substantive influence humans have had on the state, dynamics and future of the Earth System.
Anthropogenic: Resulting from or produced by human activities.
Anthropogenic emissions: See Emissions
Anthropogenic subsidence: Downward motion of the land surface induced by anthropogenic drivers (e.g., loading,
extraction of hydrocarbons and/or groundwater, drainage, mining activities) causing sediment compaction or
subsidence/deformation of the sedimentary sequence, or oxidation of organic material, thereby leading to relative
sea level rise.

3
Arid zone: Areas where vegetation growth is severely constrained due to limited water availability. For the most
part, the native vegetation of arid zones is sparse. There is high rainfall variability, with annual averages below 300
mm. Crop farming in arid zones requires irrigation.
Aridity:
The state of a long-term climatic feature characterized by low average precipitation or available water in a region.
Aridity generally arises from widespread persistent atmospheric subsidence or anticyclonic conditions, and from
more localised subsidence in the lee side of mountains (adapted from Gbeckor-Kove, 1989; Türkeş, 1999).
Atmosphere: :The gaseous envelope surrounding the earth, divided into five layers – the troposphere which
contains half of the Earth’s atmosphere, the stratosphere, the mesosphere, the thermosphere, and the exosphere,
which is the outer limit of the atmosphere. The dry atmosphere consists almost entirely of nitrogen (78.1% volume
mixing ratio) and oxygen (20.9% volume mixing ratio), together with a number of trace gases, such as argon (0.93
% volume mixing ratio), helium and radiatively active greenhouse gases (GHGs) such as carbon dioxide (CO2)
(0.04% volume mixing ratio) and ozone (O3). In addition, the atmosphere contains the GHG water vapour (H2O),
whose amounts are highly variable but typically around 1% volume mixing ratio. The atmosphere also contains
clouds and aerosols.
Attribution: See Detection and Attribution
Autonomous adaptation: See Adaptation
Avalanche: A mass of snow, ice, earth or rocks, or a mixture of these, falling down a mountainside.
Baseline scenario: see Reference scenario
Benthic: Occurring at the bottom of a body of water; related to benthos (NOAA, 2018b).
Benthos: The community of organisms living on the bottom or in sediments of a body of water (such as an ocean, a
river or a lake). The ecological zone at the bottom of a body of water, including the sediment surface and some sub-
surface layers, is known as the ‘benthic zone’.
Biodiversity: Biodiversity or biological diversity means the variability among living organisms from all sources
including, among other things, terrestrial, marine and other aquatic ecosystems, and the ecological complexes of
which they are part; this includes diversity within species, between species, and of ecosystems (UN, 1992).
Biodiversity Hotspots: Biodiversity hotspots are geographic areas exceptionally rich in species, ecologically
distinct, and often contain geographically rare endemic species. They are thus priorities for nature conservation
action.
Bioenergy: Energy derived from any form of biomass or its metabolic by-products.
Biofuel: A fuel, generally in liquid form, produced from biomass. Biofuels include bioethanol from sugarcane,
sugar beet or maize, and biodiesel from canola or soybeans.
Biomass: Organic material excluding the material that is fossilised or embedded in geological formations. Biomass
may refer to the mass of organic matter in a specific area (ISO, 2014).
Biome: Global-scale zones, generally defined by the type of plant life that they support in response to average
rainfall and temperature patterns. For example, tundra, coral reefs or savannas’ (IPBES, 2019).
Biosphere (terrestrial and marine): The part of the earth system comprising all ecosystems and living organisms,
in the atmosphere, on land (terrestrial biosphere) or in the oceans (marine biosphere), including derived dead
organic matter, such as litter, soil organic matter and oceanic detritus.
Blue carbon: All biologically-driven carbon fluxes and storage in marine systems that are amenable to
management can be considered as blue carbon. Coastal blue carbon focuses on rooted vegetation in the coastal
zone, such as tidal marshes, mangroves and seagrasses. These ecosystems have high carbon burial rates on a per
unit area basis and accumulate carbon in their soils and sediments. They provide many non-climatic benefits and
can contribute to ecosystem-based adaptation. If degraded or lost, coastal blue carbon ecosystems are likely to

4
release most of their carbon back to the atmosphere. There is current debate regarding the application of the blue
carbon concept to other coastal and non-coastal processes and ecosystems, including the open ocean.
Burden: The total mass of a gaseous substance of concern in the atmosphere.
Business as usual (BAU): :The term business as usual scenario has been used to describe a scenario that assumes
no additional policies beyond those currently in place and that patterns of socio-economic development are
consistent with recent trends. The term is now used less frequently than in the past. See also Reference scenario
Calcification: The process of biologically precipitating calcium carbonate minerals to create organism shells,
skeletons, otoliths, or other body structures. The chemical equation describing calcification is Ca2+(aq) +
2HCO3−(aq) → CaCO3(s) + CO2 + H2O. Aragonite and calcite are two common crystalline forms of biologically
precipitated calcium carbonate minerals that have different solubilities.
Capacity building: The practice of enhancing the strengths and attributes of, and resources available to, an
individual, community, society, or organization to respond to change.
Carbon dioxide (CO2): A naturally occurring gas, CO2 is also a by-product of burning fossil fuels (such as oil,
gas and coal), of burning biomass, of land use changes (LUC) and of industrial processes (e.g., cement production).
It is the principal anthropogenic greenhouse gas (GHG) that affects the Earth's radiative balance. It is the reference
gas against which other GHGs are measured and therefore has a Global Warming Potential (GWP) of 1.
Carbon dioxide (CO2) fertilization: The increase of plant photosynthesis and water-use efficiency in response to
increased atmospheric carbon dioxide (CO2) concentration. Whether this increased photosynthesis translates into
increased plant growth and carbon storage on land depends on the interacting effects of temperature, moisture and
nutrient availability.
Carbon footprint: Measure of the exclusive total amount of emissions of carbon dioxide (CO2) that is directly and
indirectly caused by an activity or is accumulated over the life stages of a product (Wiedmann and Minx, 2008).
Carbon stock: The quantity of carbon in a carbon pool.
Cascading impacts: Cascading impacts from extreme weather/climate events occur when an extreme hazard
generates a sequence of secondary events in natural and human systems that result in physical, natural, social or
economic disruption, whereby the resulting impact is significantly larger than the initial impact. Cascading impacts
are complex and multi-dimensional, and are associated more with the magnitude of vulnerability than with that of
the hazard (modified from Pescaroli & Alexander, 2015).
Catchment: An area that collects and drains precipitation.
Citizen science: A voluntary participation of the public in the collection and/or processing of data as part of a
scientific study (Silvertown, 2009).
Climate: In a narrow sense climate is usually defined as the average weather -or more rigorously, as the statistical
description in terms of the mean and variability of relevant quantities- over a period of time ranging from months to
thousands or millions of years. The classical period for averaging these variables is 30 years, as defined by the
World Meteorological Organization (WMO). The relevant quantities are most often surface variables such as
temperature, precipitation and wind. Climate in a wider sense is the state, including a statistical description, of the
climate system.
Climate change: A change in the state of the climate that can be identified (e.g., by using statistical tests) by
changes in the mean and/or the variability of its properties and that persists for an extended period, typically
decades or longer. Climate change may be due to natural internal processes or external forcings such as
modulations of the solar cycles, volcanic eruptions and persistent anthropogenic changes in the composition of the
atmosphere or in land use. Note that the United Nations Framework Convention on Climate Change (UNFCCC), in
its Article 1, defines climate change as: 'a change of climate which is attributed directly or indirectly to human
activity that alters the composition of the global atmosphere and which is in addition to natural climate variability
observed over comparable time periods'. The UNFCCC thus makes a distinction between climate change
attributable to human activities altering the atmospheric composition and climate variability attributable to natural
causes.

5
Climate extreme (extreme weather or climate event): The occurrence of a value of a weather or climate variable
above (or below) a threshold value near the upper (or lower) ends of the range of observed values of the variable.
Climate feedback: An interaction in which a perturbation in one climate quantity causes a change in a second and
the change in the second quantity ultimately leads to an additional change in the first. A negative feedback is one in
which the initial perturbation is weakened by the changes it causes; a positive feedback is one in which the initial
perturbation is enhanced. The initial perturbation can either be externally forced or arise as part of internal
variability.
Climate finance: There is no agreed definition of climate finance. The term 'climate finance' is applied both to the
financial resources devoted to addressing climate change globally and to financial flows to developing countries to
assist them in addressing climate change. The literature includes several concepts in these categories, among which
the most commonly used include:
Incremental costs: The cost of capital of the incremental investment and the change of operating and maintenance
costs for a mitigation or adaptation project in comparison to a reference project. It can be calculated as the
difference of the net present values of the two projects.
Incremental investment: The extra capital required for the initial investment for a mitigation or adaptation project
in comparison to a reference project.
Total climate finance: All financial flows whose expected effect is to reduce net greenhouse gas (GHG) emissions
and/or to enhance resilience to the impacts of climate variability and the projected climate change. This covers
private and public funds, domestic and international flows and expenditures for mitigation and adaptation to current
climate variability as well as future climate change. Total climate finance flowing to developing countries The
amount of the total climate finance invested in developing countries that comes from developed countries. This
covers private and public funds.
Private climate finance flowing to developing countries: Finance and investment by private actors in/from
developed countries for mitigation and adaptation activities in developing countries.
Public climate finance flowing to developing countries: Finance provided by developed countries' governments
and bilateral institutions as well as by multilateral institutions for mitigation and adaptation activities in developing
countries. Most of the funds provided are concessional loans and grants.
Climate governance: See Governance
Climate justice: See Justice
Climate model: A qualitative or quantitative representation of the climate system based on the physical, chemical
and biological properties of its components, their interactions and feedback processes and accounting for some of
its known properties. The climate system can be represented by models of varying complexity; that is, for any one
component or combination of components a spectrum or hierarchy of models can be identified, differing in such
aspects as the number of spatial dimensions, the extent to which physical, chemical or biological processes are
explicitly represented, or the level at which empirical parametrisations are involved.
There is an evolution towards more complex models with interactive chemistry and biology. Climate models are
applied as a research tool to study and simulate the climate and for operational purposes, including monthly,
seasonal and interannual climate predictions.
Climate prediction: A climate prediction or climate forecast is the result of an attempt to produce (starting from a
particular state of the climate system) an estimate of the actual evolution of the climate in the future, for example, at
seasonal, interannual, or decadal time scales. Because the future evolution of the climate system may be highly
sensitive to initial conditions, such predictions are usually probabilistic in nature.
Climate projection: Simulated response of the climate system to a scenario of future emissions or concentrations
of greenhouse gases (GHGs) and aerosols and changes in land use, generally derived using climate models. Climate
projections depend on an emission / concentration / radiative forcing scenario, which is in turn based on
assumptions concerning, for example, future socioeconomic and technological developments that may or may not
be realised.

6
Climate scenario: See Scenarios
Climate services: Information and products that enhance users' knowledge and understanding about the impacts of
climate change and/or climate variability so as to aid decision-making of individuals and organizations and enable
preparedness and early climate change action. Such services involve high-quality data from national and
international databases on temperature, rainfall, wind, soil moisture and ocean conditions, as well as maps, risk and
vulnerability analyses, assessments, and long-term projections and scenarios. Depending on the user’s needs, these
data and information products may be combined with non-meteorological data, such as agricultural production,
health trends, population distributions in high-risk areas, road and infrastructure maps for the delivery of goods, and
other socio-economic variables (WMO, 2019).
Climate system: The global system consisting of five major components: the atmosphere, the hydrosphere, the
cryosphere, the lithosphere and the biosphere and the interactions between them. The climate system changes in
time under the influence of its own internal dynamics and because of external forcings such as volcanic eruptions,
solar variations, orbital forcing, and anthropogenic forcings such as the changing composition of the atmosphere
and land-use change.
Climate variability: Deviations of some climate variables from a given mean state (including the occurrence of
extremes, etc.) at all spatial and temporal scales beyond that of individual weather events. Variability may be
intrinsic, due to fluctuations of processes internal to the climate system (internal variability), or extrinsic, due to
variations in natural or anthropogenic external forcing (forced variability).
Climate velocity: The speed at which isolines of a specified climate variable travel across landscapes or seascapes
due to changing climate. For example, climate velocity for temperature is the speed at which isotherms move due to
changing climate (km yr-1) and is calculated as the temporal change in temperature (°C yr-1) divided by the current
spatial gradient in temperature (°C km-1). It can be calculated using additional climate variables such as
precipitation or can be based on the climatic niche of organisms.
Climate-resilient development pathways (CRDPs): See Pathways
Climate-resilient pathways: See Pathways
Climate-smart agriculture (CSA): An approach to agriculture that aims to transform and reorient agricultural
systems to effectively support development and ensure food security in a changing climate by: sustainably
increasing agricultural productivity and incomes; adapting and building resilience to climate change; and reducing
and/or removing greenhouse gas emissions, where possible (FAO, 2018).
Climatic driver (Climate driver): A changing aspect of the climate system that influences a component of a
human or natural system.
CMIP3, CMIP5 and CMIP6 : See Coupled Model Intercomparison Project (CMIP)
Co-benefits: The positive effects that a policy or measure aimed at one objective might have on other objectives,
thereby increasing the total benefits for society or the environment. Co-benefits are often subject to uncertainty and
depend on local circumstances and implementation practices, among other factors. Co-benefits are also referred to
as ancillary benefits.
Coast : The land adjoining or near to the sea.
Coastal: The term ‘coastal’ can refer to that land (e.g., as in ‘coastal communities’), or to that part of the marine
environment that is strongly influenced by land-based processes. Thus, coastal seas are generally shallow and near-
shore. The landward and seaward limits of the coastal zone are not consistently defined, neither scientifically nor
legally. Thus, coastal waters can either be considered as equivalent to territorial waters (extending 12 nautical miles
/ 22.2 km from mean low water), or to the full Exclusive Economic Zone, or to shelf seas, with less than 200 m
water depth.
Coastal Erosion: Coastal erosion, sometimes referred to as shoreline retreat, occurs when a net loss of sediment or
bedrock from the shoreline results in landward movement of the high-tide mark.

7
Cold days/cold nights: Days where maximum temperature, or nights where minimum temperature, falls below the
10th percentile, where the respective temperature distributions are generally defined with respect to the 1961-1990
reference period. For the corresponding indices, see Box 2.4.
Communicable diseases: Illness due to a specific infectious agent or its toxic products that arises through
transmission of that agent or its products from an infected person, animal, or reservoir to a susceptible host, either
directly or indirectly through an intermediate plant or animal host, vector, or the inanimate environment."
Communicable disease pathogens include bacteria, viruses, fungi, parasites and prions.
Community-based adaptation: See Adaptation
Compound weather/climate events: The combination of multiple drivers and/or hazards that contributes to
societal and environmental risk (Zscheischler et al., 2018).
Compound risks: See Risk
Confidence: The robustness of a finding based on the type, amount, quality and consistency of evidence (e.g.,
mechanistic understanding, theory, data, models, expert judgment) and on the degree of agreement across multiple
lines of evidence. In this Report, confidence is expressed qualitatively (Mastrandrea et al., 2010.
Coping:
The use of available skills, resources, and opportunities to address, manage, and overcome adverse conditions, with
the aim of achieving basic functioning of people, institutions, organizations, and systems in the short to medium
term. FTN: This glossary entry builds from the definition used in UNISDR (2009) and IPCC (2012a).
Coral bleaching: Loss of coral pigmentation through the loss of intracellular symbiotic algae (known as
zooxanthellae) and/or loss of their pigments.
Coral reef: An underwater ecosystem characterised by structure-building stony corals. Warm-water coral reefs
occur in shallow seas, mostly in the tropics, with the corals (animals) containing algae (plants) that depend on light
and relatively stable temperature conditions. Cold-water coral reefs occur throughout the world, mostly at water
depths of 50-500 m. In both kinds of reef, living corals frequently grow on older, dead material, predominantly
made of calcium carbonate (CaCO3). Both warm and cold-water coral reefs support high biodiversity of fish and
other groups, and are considered to be especially vulnerable to climate change.
Cost-benefit analysis: Monetary assessment of all negative and positive impacts associated with a given action.
Cost-benefit analysis enables comparison of different interventions, investments or strategies and reveal how a
given investment or policy effort pays off for a particular person, company or country. Cost-benefit analyses
representing society's point of view are important for climate change decision-making, but there are difficulties in
aggregating costs and benefits across different actors and across timescales.
Coupled Model Intercomparison Project (CMIP): A climate modelling activity from the World Climate
Research Programme (WCRP) which coordinates and archives climate model simulations based on shared model
inputs by modelling groups from around the world. The CMIP3 multi-model data set includes projections using
Special Report on Emissions Scenarios (SRES) scenarios. The CMIP5 data set includes projections using the
Representative Concentration Pathways (RCP). The CMIP6 phase involves a suite of common model experiments
as well as an ensemble of CMIP-endorsed Model Intercomparison Projects (MIPs).
Cryosphere: The components of the Earth System at and below the land and ocean surface that are frozen,
including snow cover, glaciers, ice sheets, ice shelves, icebergs, sea ice, lake ice, river ice, permafrost and
seasonally frozen ground.
Cultural Impacts: Impacts on material and ecological aspects of culture and the lived experience of culture,
including dimensions such as identity, community cohesion and belonging, sense of place, worldview, values,
perceptions, and tradition. Cultural impacts are closely related to ecological impacts, especially for iconic and
representational dimensions of species and landscapes. Culture and cultural practices frame the importance and
value of the impacts of change, shape the feasibility and acceptability of adaptation options, and provide the skills
and practices that enable adaptation.

8
Decarbonisation Process by which countries, individuals or other entities aim to achieve zero fossil carbon
existence. Typically refers to a reduction of the carbon emissions associated with electricity, industry and transport.
Deep uncertainty: 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).
Deforestation: Conversion of forest to non-forest. [Note: For a discussion of the term forest and related terms such
as afforestation, reforestation and deforestation in the context of reporting and accounting Article 3.3 and 3.4
activities under the Kyoto Protocol, see 2013 Revised Supplementary Methods and Good Practice Guidance
Arising from the Kyoto Protocol.]
Desertification: Land degradation in arid, semi-arid, and dry sub-humid areas resulting from many factors,
including climatic variations and human activities (UNCCD, 1994).
Detection and attribution: Detection of change is defined as the process of demonstrating that climate or a system
affected by climate has changed in some defined statistical sense, without providing a reason for that change. An
identified change is detected in observations if its likelihood of occurrence by chance due to internal variability
alone is determined to be small, for example, <10%. Attribution is defined as the process of evaluating the relative
contributions of multiple causal factors to a change or event with a formal assessment of confidence.
Developed / developing countries (Industrialised / developed / developing countries): There is a diversity of
approaches for categorizing countries on the basis of their level of development, and for defining terms such as
industrialised, developed, or developing. Several categorisations are used in this Special Report. (1) In the United
Nations (UN) system, there is no established convention for the designation of developed and developing countries
or areas. (2) The UN Statistics Division specifies developed and developing regions based on common practice. In
addition, specific countries are designated as least developed countries, landlocked developing countries, small
island developing states (SIDS), and transition economies. Many countries appear in more than one of these
categories. (3) The World Bank uses income as the main criterion for classifying countries as low, lower middle,
upper middle, and high income. (4) The UN Development Programme (UNDP) aggregates indicators for life
expectancy, educational attainment, and income into a single composite Human Development Index (HDI) to
classify countries as low, medium, high, or very high human development.
Development pathways: See Pathways
Diatoms: Silt-sized algae that live in surface waters of lakes, rivers and oceans and form shells of opal. Their
species distribution in ocean cores is often related to past sea surface temperatures.
Disaster: A ‘serious disruption of the functioning of a community or a society at any scale due to hazardous events
interacting with conditions of exposure, vulnerability and capacity, leading to one or more of the following: human,
material, economic and environmental losses and impacts’ (UNISDR, 2017).
Disaster management: Social processes for designing, implementing, and evaluating strategies, policies, and
measures that promote and improve disaster preparedness, response, and recovery practices at different
organizational and societal levels.
Disaster risk: The likelihood over a specified time period of severe alterations in the normal functioning of a
community or a society due to hazardous physical events interacting with vulnerable social conditions, leading to
widespread adverse human, material, economic, or environmental effects that require immediate emergency
response to satisfy critical human needs and that may require external support for recovery.
Disaster risk management (DRM): Processes for designing, implementing, and evaluating strategies, policies,
and measures to improve the understanding of current and future disaster risk, foster disaster risk reduction and
transfer, and promote continuous improvement in disaster preparedness, prevention and protection, response, and
recovery practices, with the explicit purpose of increasing human security, wellbeing, quality of life, and
sustainable development (SD).

9
Disaster Risk Reduction (DRR): Denotes both a policy goal or objective, and the strategic and instrumental
measures employed for anticipating future disaster risk; reducing existing exposure, hazard, or vulnerability; and
improving resilience.
Discount rate: See Discounting
Discounting : A mathematical operation that aims to make monetary (or other) amounts received or expended at
different times (years) comparable across time. The discounter uses a fixed or possibly time-varying discount rate
from year to year that makes future value worth less today (if the discount rate is positive). The choice of discount
rate(s) is debated as it is a judgement based on hidden and/or explicit values.
Displacement: In land system science, displacement denotes the increasing spatial separation between the location
of agricultural and forestry production and the place of consumption of these products, as it occurs with trade.
Displacement disconnects spatially environmental impacts from their socioeconomic drivers.
Downscaling: A method that derives local- to regional-scale (up to 100 km) information from larger-scale models
or data analyses. Two main methods exist: dynamical downscaling and empirical/statistical downscaling. The
dynamical method uses the output of regional climate models, global models with variable spatial resolution, or
high-resolution global models. The empirical/statistical methods are based on observations and develop statistical
relationships that link the large-scale atmospheric variables with local/regional climate variables. In all cases, the
quality of the driving model remains an important limitation on quality of the downscaled information. The two
methods can be combined, e.g., applying empirical/statistical downscaling to the output of a regional climate
model, consisting of a dynamical downscaling of a global climate model.
Drainage: Artificial lowering of the soil water table (IPCC, 2013).
Driver: Any natural or human-induced factor that directly or indirectly causes a change in a system (adapted from
MA, 2005). See also Climate-driver
Drought: A period of abnormally dry weather long enough to cause a serious hydrological imbalance. Drought is a
relative term; therefore, any discussion in terms of precipitation deficit must refer to the particular precipitation-
related activity that is under discussion. For example, shortage of precipitation during the growing season impinges
on crop production or ecosystem function in general (due to soil moisture drought, also termed agricultural drought)
and during the runoff and percolation season primarily affects water supplies (hydrological drought). Storage
changes in soil moisture and groundwater are also affected by increases in actual evapotranspiration in addition to
reductions in precipitation. A period with an abnormal precipitation deficit is defined as a meteorological drought.
Megadrought: A very lengthy and pervasive drought, lasting much longer than normal, usually a decade or more.
Dynamic Vulnerability: Describes the quality of social and physical vulnerability to change between consecutive
hazard exposures (or indeed other stresses and shocks such as health – i.e. pandemic; economic i.e. price rises or
political i.e. violence). This emphasises the need for assessments of risk to also be dynamic and able to capture
change over time. See also cascading impacts
Early warning systems (EWS): The set of technical and institutional capacities to forecast, predict, and
communicate timely and meaningful warning information to enable individuals, communities, managed
ecosystems, and organisations threatened by a hazard to prepare to act promptly and appropriately to reduce the
possibility of harm or loss. Dependent upon context, EWS may draw upon scientific and/or indigenous knowledge,
and other knowledge types. EWS are also considered for ecological applications, e.g., conservation, where the
organisation itself is not threatened by hazard but the ecosystem under conservation is (e.g., coral bleaching alerts),
in agriculture (e.g., warnings of heavy rainfall, drought, ground frost, and hailstorms) and in fisheries (e.g.,
warnings of storm, storm surge, and tsunamis) (UNISDR 2009; IPCC, 2012a).
Earth system model (ESM): A coupled atmosphere–ocean general circulation model (AOGCM) in which a
representation of the carbon cycle is included, allowing for interactive calculation of atmospheric carbon dioxide
(CO2) or compatible emissions. Additional components (e.g., atmospheric chemistry, ice sheets, dynamic
vegetation, nitrogen cycle, but also urban or crop models) may be included.

10
Eastern boundary upwelling system (EBUS): Eastern boundary upwelling system (EBUS) are located at the
eastern (landward) edges of major ocean basins in both hemispheres, where equatorward winds drive upwelling
currents that bring cool, nutrient-rich (and often oxygen-poor) waters from the deep ocean to the surface near the
coast.
Ecosystem: A functional unit consisting of living organisms, their non-living environment and the interactions
within and between them. The components included in a given ecosystem and its spatial boundaries depend on the
purpose for which the ecosystem is defined: in some cases, they are relatively sharp, while in others they are
diffuse. Ecosystem boundaries can change over time. Ecosystems are nested within other ecosystems and their scale
can range from very small to the entire biosphere. In the current era, most ecosystems either contain people as key
organisms, or are influenced by the effects of human activities in their environment.
Ecosystem services: Ecological processes or functions having monetary or non-monetary value to individuals or
society at large. These are frequently classified as (1) supporting services such as productivity or biodiversity
maintenance, (2) provisioning services such as food or fibre, (3) regulating services such as climate regulation or
carbon sequestration, and (4) cultural services such as tourism or spiritual and aesthetic appreciation.
Ecosystem-based adaptation (EBA): See Adaptation
Efficacy: A measure of how effective a radiative forcing from a given anthropogenic or natural mechanism is at
changing the equilibrium global mean surface temperature compared to an equivalent radiative forcing from carbon
dioxide. A carbon dioxide increase by definition has an efficacy of 1.0. Variations in climate efficacy may result
from rapid adjustments to the applied forcing, which differ with different forcings.
El Niño-Southern Oscillation (ENSO): The term El Niño was initially used to describe a warm-water current that
periodically flows along the coast of Ecuador and Peru, disrupting the local fishery. It has since become identified
with warming of the tropical Pacific Ocean east of the dateline. This oceanic event is associated with a fluctuation
of a global-scale tropical and subtropical surface pressure pattern called the Southern Oscillation. This coupled
atmosphere– ocean phenomenon, with preferred time scales of two to about seven years, is known as the El Niño-
Southern Oscillation (ENSO). It is often measured by the surface pressure anomaly difference between Tahiti and
Darwin and/or the sea surface temperatures in the central and eastern equatorial Pacific. During an ENSO event, the
prevailing trade winds weaken, reducing upwelling and altering ocean currents such that the sea surface
temperatures warm, further weakening the trade winds. This phenomenon has a great impact on the wind, sea
surface temperature and precipitation patterns in the tropical Pacific. It has climatic effects throughout the Pacific
region and in many other parts of the world, through global teleconnections. The cold phase of ENSO is called La
Niña.
Emergence (of the Climate Signal): When a change in climate (the ‘signal’) becomes larger than the amplitude of
natural or internal variations (defining the ‘noise’), it is termed to have ‘emerged’. This concept is often expressed
as a ‘signal-to-noise’ ratio and emergence occurs at a defined threshold of this ratio. Emergence can refer to
changes relative to a historical or modern baseline, and can be expressed in terms of time or in terms of a global
warming level. Emergence can be estimated using observations and/or model simulations.
Emission pathways: See Pathways
Emission scenario: See Scenario
Emissions: Anthropogenic emissions: Emissions of greenhouse gases (GHGs), precursors of GHGs and aerosols
caused by human activities. These activities include the burning of fossil fuels, deforestation, land use and land use
changes (LULUC), livestock production, fertilisation, waste management, and industrial processes. Fossil fuel
emissions: Emissions of greenhouse gases (in particular carbon dioxide), other trace gases and aerosols resulting
from the combustion of fuels from fossil carbon deposits such as oil, gas and coal. Non-CO2 emissions and
radiative forcing: Non-CO2 emissions included in this report are all anthropogenic emissions other than CO2 that
result in radiative forcing. These include short-lived climate forcers, such as methane (CH4), some fluorinated
gases, ozone (O3) precursors, aerosols or aerosol precursors, such as black carbon and sulphur dioxide,
respectively, as well as long-lived greenhouse gases, such as nitrous oxide (N2O) or other fluorinated gases. The
radiative forcing associated with non-CO2 emissions and changes in surface albedo is referred to as non-CO2
radiative forcing.

11
Enabling conditions (for adaptation and mitigation options): Conditions that affect the feasibility of adaptation
and mitigation options, and can accelerate and scale-up systemic transitions that would limit temperature increase
and enhance capacities of systems and societies to adapt to the associated climate change, while achieving
sustainable development, eradicating poverty and reducing inequalities. Enabling conditions include finance,
technological innovation, strengthening policy instruments, institutional capacity, multi-level governance, and
changes in human behaviour and lifestyles. They also include inclusive processes, attention to power asymmetries
and unequal opportunities for development and reconsideration of values.
Endemic species: Plants and animals that are only found in one geographic region.
Energy access: Access to clean, reliable and affordable energy services for cooking and heating, lighting,
communications, and productive uses (with special reference to Sustainable Development Goal 7) (AGECC, 2010).
Energy efficiency: The ratio of output or useful energy or energy services or other useful physical outputs obtained
from a system, conversion process, transmission or storage activity to the input of energy (measured as kWh kWh-
1, tonnes kWh-1 or any other physical measure of useful output like tonne-km transported). Energy efficiency is
often described by energy intensity. In economics, energy intensity describes the ratio of economic output to energy
input. Most commonly energy efficiency is measured as input energy over a physical or economic unit, i.e., kWh
USD-1 (energy intensity), kWh tonne-1. For buildings, it is often measured as kWh m-2, and for vehicles as km
liter-1 or liter km-1. Very often in policy ‘energy efficiency’ is intended as the measures to reduce energy demand
through technological options such as insulating buildings, more efficient appliances, efficient lighting, efficient
vehicles, etc.
Energy security: The goal of a given country, or the global community as a whole, to maintain an adequate, stable
and predictable energy supply. Measures encompass safeguarding the sufficiency of energy resources to meet
national energy demand at competitive and stable prices and the resilience of the energy supply; enabling
development and deployment of technologies; building sufficient infrastructure to generate, store and transmit
energy supplies and ensuring enforceable contracts of delivery.
Energy system: The energy system comprises all components related to the production, conversion, delivery, and
use of energy.
Equality: A principle that ascribes equal worth to all human beings, including equal opportunities, rights, and
obligations, irrespective of origins.
Inequality: Uneven opportunities and social positions, and processes of discrimination within a group or society,
based on gender, class, ethnicity, age, and (dis)ability, often produced by uneven development. Income inequality
refers to gaps between highest and lowest income earners within a country and between countries.
Equity: The principle of being fair and impartial, and a basis for understanding how the impacts and responses to
climate change, including costs and benefits, are distributed in and by society in more or less equal ways. Often
aligned with ideas of equality, fairness and justice and applied with respect to equity in the responsibility for, and
distribution of, climate impacts and policies across society, generations, and gender, and in the sense of who
participates and controls the processes of decision-making.
Ethics: Ethics involves questions of justice and value. Justice is concerned with right and wrong, equity and
fairness, and, in general, with the rights to which people and living beings are entitled. Value is a matter of worth,
benefit, or good.
Eutrophication: Over-enrichment of water by nutrients such as nitrogen and phosphorus. It is one of the leading
causes of water quality impairment. The two most acute symptoms of eutrophication are hypoxia (or oxygen
depletion) and harmful algal blooms.
Evaporation: The physical process by which a liquid (e.g., water) becomes a gas (e.g., water vapour).

12
Evapotranspiration The combined processes through which water is transferred to the atmosphere from open
water and ice surfaces, bare soil, and vegetation that make up the Earth’s surface.
Evidence: Data and information used in the scientific process to establish findings. In this report, the degree of
evidence reflects the amount, quality and consistency of scientific/technical information on which the Lead Authors
are basing their findings.
Evolutionary adaptation: See Adaptation
Exposure: The presence of people; livelihoods; species or ecosystems; environmental functions, services, and
resources; infrastructure; or economic, social, or cultural assets in places and settings that could be adversely
affected.
Externality/external cost/external benefit: Externalities arise from a human activity, when agents responsible for
the activity do not take full account of the activity's impact on others' production and consumption possibilities, and
no compensation exists for such impacts. When the impact is negative, they are external costs. When positive they
are referred to as external benefits.
Extinction: A population, species or more inclusive taxonomic group has gone extinct when all its individuals have
died. A species may go extinct locally (population extinction), regionally (e.g., extinction of all populations in a
country, continent or ocean) or globally (IPBES, 2019. Global Assessment).
Extreme/ heavy precipitation event: An extreme/heavy precipitation event is an event that is of very high
magnitude with a 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 days), though all of them qualitatively represent high magnitude.
The intensity of such events may be defined with block maxima approach such as annual maxima or with peak over
threshold approach, such as rainfall above 95th or 99th percentile at a particular space.
Extreme sea level: Extreme sea levels (ESLs) arise from a combination of surges, tides, and waves, superimposed
on the local sea-surface height. The combination of tide and surge is referred to as still water level and the
combination of tide, surge, and wave effects is referred to as total water level. The wave contribution (wave run-up,
the maximum height waves reach the beach) can be characterized as the combination of the wave setup (time- mean
sea-level elevation due to wave energy dissipation) and the swash (vertical displacement induced by individual
waves). RSL change affects ESL directly by shifting the base water levels and indirectly by modulating propagation
of tides, waves, and/or surges due to increased water depth. In addition, ESLs may change due to changes in the
frequency, tracks, or strength of weather systems, or due to anthropogenically induced changes such as the
modification of coastlines.
Extreme weather event: An event that is rare at a particular place and time of year. Definitions of ‘rare’ vary, but
an extreme weather event would normally be as rare as or rarer than the 10th or 90th percentile of a probability
density function estimated from observations. By definition, the characteristics of what is called extreme weather
may vary from place to place in an absolute sense. When a pattern of extreme weather persists for some time, such
as a season, it may be classed as an extreme climate event, especially if it yields an average or total that is itself
extreme (e.g., drought or heavy rainfall over a season).
Feasibility: The degree to which climate goals and response options are considered possible and/or desirable.
Feasibility depends on geophysical, ecological, technological, economic, social and institutional conditions for
change. Conditions underpinning feasibility are dynamic, spatially variable, and may vary between different
groups.
Flood: The overflowing of the normal confines of a stream or other water body, or the accumulation of water over
areas that are not normally submerged. Floods can be caused by unusually heavy rain, for example during storms
and cyclones. Floods include river (fluvial) floods, flash floods, urban floods, rain (pluvial) floods, sewer floods,
coastal floods, and glacial lake outburst floods (GLOFs).
Flux: A movement (a flow) of matter (e.g., water vapor, particles), heat or energy from one place to another, or
from one medium (e.g., land surface) to another (e.g., atmosphere).

13
Food-borne disease: Illnesses transmitted through consumption of unsafe or contaminated food, that
contamination can come from a variety of sources, including contaminated water. (adapted from UNEP, 2018)
Food security: A situation that exists when all people, at all times, have physical, social and economic access to
sufficient, safe and nutritious food that meets their dietary needs and food preferences for an active and healthy life.
The four pillars of food security are availability, access, utilization and stability. The nutritional dimension is
integral to the concept of food security {FAO, 2018/ 2009 #6619}.
Availability: Physical availability of food. Food availability addresses the supply side of food security and is
determined by the levels of food production, stocks and net trade.
Access: Economic and/ or physical access to food. Economic access is determined by disposable income, food
prices and the provision of and access to social support. Physical access is determined by the availability
and quality of land and other infrastructure, property rights or the functioning of markets.
Utilisation: The way in which the body uses the various nutrients in food. Individuals achieve sufficient energy and
nutrient intake through good care and feeding practices, food preparation, diet diversity and intrahousehold
distribution of food. Combined with biological utilization of the food consumed, energy and nutrient intake
determine the nutrition status of individuals.
Stability: The stability of the other three dimensions over time. Even if individuals’ food intake is adequate today,
they are still considered food-insecure if periodically they have inadequate access to food, risking deterioration of
their nutrition status. Adverse weather conditions, political instability or economic factors (unemployment, rising
food prices) may have an impact on individuals’ food security status.
Food system: All the elements (environment, people, inputs, processes, infrastructures, institutions, etc.) and
activities that relate to the production, processing, distribution, preparation and consumption of food, and the output
of these activities, including socio-economic and environmental outcomes (HLPE, 2017). [Note: Whilst there is a
global food system (encompassing the totality of global production and consumption), each location’s food system
is unique, being defined by that place’s mix of food produced locally, nationally, regionally or globally.]
Forest: A vegetation type dominated by trees. Many definitions of the term forest are in use throughout the world,
reflecting wide differences in biogeophysical conditions, social structure and economics. [Note: For a discussion of
the term forest in the context of National GHG inventories, see the 2006 IPCC Guidelines for National GHG
Inventories and information provided by the United Nations Framework Convention on Climate Change
(UNFCCC, 2019).]
Forest degradation: A reduction in the capacity of a forest to produce ecosystem services such as carbon storage
and wood products as a result of anthropogenic and environmental changes.
Forest Dieback: See Forest degradation
Forest line: The upper limit of the closed upper montane forest or forest at high latitudes. It is less elevated or less
poleward than the tree line
Forest management: A system of practices for stewardship and use of forest land aimed at fulfilling relevant
ecological (including biological diversity), economic and social functions of the forest in a sustainable manner
(UNFCCC, 2002).
Fossil fuels: Carbon-based fuels from fossil hydrocarbon deposits, including coal, oil, and natural gas.
Geoengineering: Geoengineering A broad set of methods and technologies that aim to deliberately alter the climate
system in order to alleviate the impacts of climate change. Most, but not all, methods seek to either (1) reduce the
amount of absorbed solar energy in the climate system (solar radiation management, or solar radiation modification,
SRM) or (2) increase net carbon sinks from the atmosphere at a scale sufficiently large to alter climate (i.e., carbon
dioxide removal, CDR). Scale and intent are of central importance. Two key characteristics of geoengineering
methods of particular concern are that they use or affect the climate system (e.g., atmosphere, land, or ocean)
globally or regionally and/or could have substantive unintended effects that cross national boundaries.

14
Geoengineering is different from weather modification and ecological engineering, but the boundary can be unclear
(IPCC, 2012b, p. 2).
Glacial lake outburst flood (GLOF) / Glacier lake outburst: A sudden release of water from a glacier lake,
including any of the following types – a glacier-dammed lake, a pro-glacial moraine-dammed lake or water that
was stored within, under or on the glacier.
Glacier: A perennial mass of ice, and possibly firm and snow, originating on the land surface by accumulation and
compaction of snow and showing evidence of past or present flow. A glacier typically gains mass by accumulation
of snow, and loses mass by ablation. Land ice masses of continental size (>50,000 km2) are referred to as ice sheets
(Cogley et al., 2011).
Global change: A generic term to describe global scale changes in systems, including the climate system,
ecosystems, and social-ecological systems.
Global mean sea level change: Global mean sea level (GMSL) change is the increase or decrease in the volume of
the ocean divided by the Ocean surface area. It is driven by changes in ocean density through temperature changes
(global mean thermosteric sea level change) and changes in the ocean mass as a result of changes in the cryosphere
or terrestrial water storage (barystatic sea level change).
Global mean surface air temperature (GSAT): Global average of near-surface air temperatures over land and
oceans. Changes in GSAT are often used as a measure of global temperature change in climate models but are not
observed directly. See also Global mean surface temperature (GMST).
Global mean surface temperature (GMST): Estimated global average of near-surface air temperatures over land
and sea ice, and sea surface temperature (SST) over ice-free ocean regions, with changes normally expressed as
departures from a value over a specified reference period. When estimating changes in GMST, near-surface air
temperatures over both land and oceans are also used. See also Global mean surface air temperature (GSAT).
Global warming: An increase in global mean surface temperature (GMST) averaged over a 30-year period, or the
30-year period centred on a particular year or decade, expressed relative to pre-industrial levels unless otherwise
specified. For 30-year periods that span past and future years, the current multi-decadal warming trend is assumed
to continue.
Governance: The structures, processes, and actions through which private and public actors interact to address
societal goals. This includes formal and informal institutions and the associated norms, rules, laws and procedures
for deciding, managing, implementing and monitoring policies and measures at any geographic or political scale,
from global to local.
Adaptive governance: Adjusting to changing conditions, such as climate change, through governance interactions
that seek to maintain a desired state in a social-ecological system.
Climate governance: The structures, processes, and actions through which private and public actors seek to
mitigate and adapt to climate change.
Multilevel governance: The dispersion of governance across multiple levels of jurisdiction and decision-making,
including, global, regional, national and local, as well as trans-regional and trans-national levels.
Polycentric governance: Polycentric governance involves multiple centres of decision-making with overlapping
jurisdictions. While the centres have some degree of autonomy, they also take each other into account, coordinating
their actions and seeking to resolve conflicts (Carlisle and Gruby, 2017; Jordan et al., 2018; McGinnis and Ostrom,
2012).
Governance capacity: The ability of governance institutions, leaders, and non-state and civil society to plan, co-
ordinate, fund, implement, evaluate and adjust policies and measures over the short, medium and long term,
adjusting for uncertainty, rapid change and wide-ranging impacts and multiple actors and demands.
Green Climate Fund (GCF): The Green Climate Fund was established by the 16th Session of the Conference of
the Parties (COP) in 2010 as an operating entity of the financial mechanism of the United Nations Framework
Convention on Climate Change (UNFCCC), in accordance with Article 11 of the Convention, to support projects,

15
programmes and policies and other activities in developing country Parties. The Fund is governed by a Board and
will receive guidance of the COP. The Fund is headquartered in Songdo, Republic of Korea.
Green infrastructure: The strategically planned interconnected set of natural and constructed ecological systems,
green spaces and other landscape features that can provide functions and services including air and water
purification, temperature management, floodwater management and coastal defence often with co-benefits for
human and ecological well-being. Green infrastructure includes planted and remnant native vegetation, soils,
wetlands, parks and green open spaces, as well as building and street level design interventions that incorporate
vegetation. (after Culwick and Bobbins, 2016)
Greenhouse gases (GHG): Gaseous constituents of the atmosphere, both natural and anthropogenic, that absorb
and emit radiation at specific wavelengths within the spectrum of radiation emitted by the Earth's ocean and land
surface, by the atmosphere itself, and by clouds. This property causes the greenhouse effect. Water vapour (H2O),
carbon dioxide (CO2), nitrous oxide (N2O), methane (CH4) and ozone (O3) are the primary GHGs in the Earth's
atmosphere. Human-made GHGs include sulphur hexafluoride (SF6), hydrofluorocarbons (HFCs),
chlorofluorocarbons (CFCs) and perfluorocarbons (PFCs); several of these are also O3-depleting (and are regulated
under the Montreal Protocol).
Gross domestic product (GDP): The sum of gross value added, at purchasers' prices, by all resident and non-
resident producers in the economy, plus any taxes and minus any subsidies not included in the value of the products
in a country or a geographic region for a given period, normally one year. GDP is calculated without deducting for
depreciation of fabricated assets or depletion and degradation of natural resources.
Groundwater recharge: The process by which external water is added to the zone of saturation of an aquifer,
either directly into a geologic formation that traps the water or indirectly by way of another formation.
Habitability (human): The ability of a place to support human life by providing protection from hazards which
challenge human survival, and by assuring adequate space, food and freshwater.
Hazard: The potential occurrence of a natural or human-induced physical event or trend that may cause loss of life,
injury, or other health impacts, as well as damage and loss to property, infrastructure, livelihoods, service provision,
ecosystems and environmental resources.
Health: Health is a state of complete physical, mental and social well-being and not merely the absence of disease
or firmity (WHO).
Heat stress: A range of conditions caused by being over exposed to high temperatures, which exert serious impacts
on particular exposed objects or systems. The risks of heat stress are heterogeneous due to differences in exposure
and vulnerability for various regions and objects.
Heat wave : A period of abnormally hot weather. Heatwaves and warm spells have various and, in some cases,
overlapping definitions. (see also Marine Heatwave)
Heavy precipitation event: See Extreme/heavy precipitation event.
Human mobility: The permanent or semi-permanent move by a person for at least one year and involving crossing
an administrative, but not necessarily a national, border.
Human rights: Rights that are inherent to all human beings, universal, inalienable, and indivisible, typically
expressed and guaranteed by law. They include the right to life, economic, social, and cultural rights, and the right
to development and self-determination (UNOHCHR, 2018).
Human security: A condition that is met when the vital core of human lives is protected, and when people have the
freedom and capacity to live with dignity. In the context of climate change, the vital core of human lives includes
the universal and culturally specific, material and non-material elements necessary for people to act on behalf of
their interests and to live with dignity.
Human system: Any system in which human organisations and institutions play a major role. Often, but not
always, the term is synonymous with society or social system. Systems such as agricultural systems, urban systems,
political systems, technological systems, and economic systems are all human systems in the sense applied in this
report.

16
Hydrological cycle: The cycle in which water evaporates from the ocean and the land surface, is carried over the
Earth in atmospheric circulation as water vapour, condenses to form clouds, precipitates over the ocean and land as
rain or snow, which on land can be intercepted by trees and vegetation, potentially accumulating as snow or ice,
provides runoff on the land surface, infiltrates into soils, recharges groundwater, discharges into streams, and
ultimately, flows into the oceans as rivers, polar glaciers and ice sheets, from which it will eventually evaporate
again. The various systems involved in the hydrological cycle are usually referred to as hydrological systems.
Hydropower: The energy of water moving from higher to lower elevations that is converted into mechanical
energy through a turbine or other device that is either used directly for mechanical work or more commonly to
operate a generator that produces electricity. The term is also used to describe the kinetic energy of stream flow that
may also be converted into mechanical energy of a generator through an in-stream turbine to produce electricity.
Hypoxic: Conditions of low dissolved oxygen in shallow water ocean and freshwater environments. There is no
universal threshold for hypoxia. A value around 60 μmol kg-1 has commonly been used for some estuarine
systems, although this does not necessarily directly translate into biological impacts. Anoxic conditions occur
where there is no oxygen present at all. See also Eutrophication.
Ice sheet: An ice body originating on land that covers an area of continental size, generally defined as covering
>50,000 km2, and that has formed over thousands of years through accumulation and compaction of snow. An ice
sheet flows outward from a high central ice plateau with a small average surface slope. The margins usually slope
more steeply, and most ice is discharged through fast-flowing ice streams or outlet glaciers, often into the sea or
into ice shelves floating on the sea. There are only two ice sheets in the modern world, one on Greenland and one
on Antarctica. The latter is divided into the East Antarctic Ice Sheet (EAIS), the West Antarctic Ice Sheet (WAIS)
and the Antarctic Peninsula ice sheet. During glacial periods, there were other ice sheets.
Impacts (consequences, outcomes): The consequences of realized risks on natural and human systems, where
risks result from the interactions of climate-related hazards (including extreme weather / climate events), exposure,
and vulnerability. Impacts generally refer to effects on lives, livelihoods, health and wellbeing, ecosystems and
species, economic, social and cultural assets, services (including ecosystem services), and infrastructure. Impacts
may be referred to as consequences or outcomes, and can be adverse or beneficial.
Income: The maximum amount that a household, or other unit, can consume without reducing its real net worth.
Total income is the broadest measure of income and refers to regular receipts such as wages and salaries, income
from self-employment, interest and dividends from invested funds, pensions or other benefits from social insurance,
and other current transfers receivable. FTN: This glossary entry builds from the definition used in OECD (2003).
Incremental adaptation: See Adaptation
Indigenous knowledge: The understandings, skills and philosophies developed by societies with long histories of
interaction with their natural surroundings. For many indigenous peoples, IK informs decision-making about
fundamental aspects of life, from day-to-day activities to longer term actions. This knowledge is integral to cultural
complexes, which also encompass language, systems of classification, resource use practices, social interactions,
values, ritual and spirituality. These distinctive ways of knowing are important facets of the world’s cultural
diversity (UNESCO, 2018).
Indigenous Peoples: Indigenous Peoples and Nations are those that, having a historical continuity with pre-
invasion and pre- colonial societies that developed on their territories, consider themselves distinct from other
sectors of the societies now prevailing on those territories, or parts of them. They form at present principally non-
dominant sectors of society and are often determined to preserve, develop, and transmit to future generations their
ancestral territories, and their ethnic identity, as the basis of their continued existence as peoples, in accordance
with their own cultural patterns, social institutions, and common law system. FTN: This glossary entry builds from
the definitions used in Cobo (1987) and previous IPCC reports.
Inequality: See Equality
Informal settlement: A term given to settlements or residential areas that by at least one criterion fall outside
official rules and regulations. Most informal settlements have poor housing (with widespread use of temporary
materials) and are developed on land that is occupied illegally with high levels of overcrowding. In most such

17
settlements, provision for safe water, sanitation, drainage, paved roads, and basic services is inadequate or lacking.
The term slum is often used for informal settlements, although it is misleading as many informal settlements
develop into good quality residential areas, especially where governments support such development.
Institutional capacity: Building and strengthening individual organisations and providing technical and
management training to support integrated planning and decision-making processes between organisations and
people, as well as empowerment, social capital, and an enabling environment, including the culture, values and
power relations (Willems and Baumert, 2003).
Institutions: Rules, norms and conventions that guide, constrain or enable human behaviours and practices.
Institutions can be formally established, for instance through laws and regulations, or informally established, for
instance by traditions or customs. Institutions may spur, hinder, strengthen, weaken or distort the emergence,
adoption and implementation of climate action and climate governance.
[Note: Institutions can also refer to a large organisation]
Insurance/reinsurance: A family of financial instruments for sharing and transferring risk among a pool of at-risk
households, businesses, and/or governments.
Integrated assessment: A method of analysis that combines results and models from the physical, biological,
economic and social sciences and the interactions among these components in a consistent framework to evaluate
the status and the consequences of environmental change and the policy responses to it.
Integrated assessment model (IAM): Models that integrate knowledge from two or more domains into a single
framework. They are one of the main tools for undertaking integrated assessments. One class of IAM used in
respect of climate change mitigation may include representations of: multiple sectors of the economy, such as
energy, land use and land use change; interactions between sectors; the economy as a whole; associated greenhouse
gas (GHG) emissions and sinks; and reduced representations of the climate system. This class of model is used to
assess linkages between economic, social and technological development and the evolution of the climate system.
Another class of IAM additionally includes representations of the costs associated with climate change impacts, but
includes less detailed representations of economic systems. These can be used to assess impacts and mitigation in a
cost-benefit framework and have been used to estimate the social cost of carbon.
Invasive species: A species that is not native to a specific location or nearby, lacking natural controls, and has a
tendency to rapidly increase in abundance, displacing native species. Invasive species may also damage the human
economy or human health.
Justice: Justice is concerned with ensuring that people get what is due to them, setting out the moral or legal
principles of fairness and equity in the way people are treated, often based on the ethics and values of society.
Climate justice: Justice that links development and human rights to achieve a human-centred approach to
addressing climate change, safeguarding the rights of the most vulnerable people and sharing the burdens and
benefits of climate change and its impacts equitably and fairly (MRFJC, 2018).
Procedural justice: Justice in the way outcomes are brought about including who participates and is heard in the
processes of decision-making.
Social justice: Just or fair relations within society that seek to address the distribution of wealth, access to
resources, opportunity, and support according to principles of justice and fairness.
Key risk: Those risks that are especially relevant to the interpretation of ‘dangerous anthropogenic interference
with the climate system’ (DAI) in the terminology of UNFCCC, Article 2, meriting particular attention by policy
makers in that context. Key risks are potentially severe adverse consequences for humans and social- ecological
systems resulting from the interaction of climate related hazards with vulnerabilities of societies and systems
exposed. Risks are considered “key” due to high hazard or high vulnerability of societies and systems exposed, or
both.
Land: The terrestrial portion of the biosphere that comprises the natural resources (soil, near-surface air, vegetation
and other biota, and water), the ecological processes, topography, and human settlements and infrastructure that
operate within that system (FAO, 2007; UNCCD, 1994).

18
Land cover:
Land cover change: Change from one land cover class to another, due to change in land use or change in natural
conditions (Pongratz et al., 2018).
Land degradation: A negative trend in land condition, caused by direct or indirect human-induced processes
including anthropogenic climate change, expressed as long-term reduction or loss of at least one of the following:
biological productivity, ecological integrity or value to humans. [Note: This definition applies to forest and non-
forest land. Changes in land condition resulting solely from natural processes (such as volcanic eruptions) are not
considered to be land degradation. Reduction of biological productivity or ecological integrity or value to humans
can constitute degradation, but any one of these changes need not necessarily be considered degradation.]
Land management: Sum of land-use practices (e.g., sowing, fertilizing, weeding, harvesting, thinning, clear-
cutting) that take place within broader land-use categories. (Pongratz et al., 2018)
Land use: The total of arrangements, activities and inputs applied to a parcel of land. The term land use is also
used in the sense of the social and economic purposes for which land is managed (e.g., grazing, timber extraction,
conservation and city dwelling). In national GHG inventories, land use is classified according to the IPCC land use
categories of forest land, cropland, grassland, wetlands, settlements, other lands (see the 2006 IPCC Guidelines for
National GHG Inventories for details).
Land-use change: The change from one land use category to another. Note that in some scientific literature, land-
use change encompasses changes in land-1 use categories as well as changes in land management. See also
Afforestation, Agriculture, Forestry and Other Land Use (AFOLU), Deforestation, Land use, land-use change and
forestry (LULUCF) and Reforestation.
Indirect land-use change (iLUC): Land use change outside the area of focus, that occurs as a consequence of
change in use or management of land within the area of focus, such as through market or policy drivers. For
example, if agricultural land is diverted to biofuel production, forest clearance may occur elsewhere to replace the
former agricultural production. See Land-use change (LUC).
Least Developed Countries (LDCs) : A list of countries designated by the Economic and Social Council of the
United Nations (ECOSOC) as meeting three criteria: (1) a low income criterion below a certain threshold of gross
national income per capita of between USD 750 and USD 900, (2) a human resource weakness based on indicators
of health, education, adult literacy, and (3) an economic vulnerability weakness based on indicators on instability of
agricultural production, instability of export of goods and services, economic importance of non-traditional
activities, merchandise export concentration, and the handicap of economic smallness. Countries in this category
are eligible for a number of programmes focused on assisting countries most in need. These privileges include
certain benefits under the articles of the United Nations Framework Convention on Climate Change (UNFCCC).
Likelihood: The chance of a specific outcome occurring, where this might be estimated probabilistically.
Likelihood is expressed in this Special Report using a standard terminology (Mastrandrea et al., 2010).
Livelihood: The resources used and the activities undertaken in order for people to live. Livelihoods are usually
determined by the entitlements and assets to which people have access. Such assets can be categorised as human,
social, natural, physical, or financial.
Local knowledge (LK): The understandings and skills developed by individuals and populations, specific to the
places where they live. Local knowledge informs decision-making about fundamental aspects of life, from day-to-
day activities to longer term actions. This knowledge is a key element of the social and cultural systems which
influence observations of and responses to climate change; it also informs governance decisions (UNESCO, 2018).
Lock-in: A situation in which the future development of a system, including infrastructure, technologies,
investments, institutions, and behavioural norms, is determined or constrained (‘locked in’) by historic
developments.
Loss and Damage, and losses and damages: Research has taken Loss and Damage (capitalised letters) to refer to
political debate under the United Nations Framework Convention on Climate Change (UNFCCC) following the
establishment of the Warsaw Mechanism on Loss and Damage in 2013, which is to ‘address loss and damage

19
associated with impacts of climate change, including extreme events and slow onset events, in developing countries
that are particularly vulnerable to the adverse effects of climate change.’ Lowercase letters (losses and damages)
have been taken to refer broadly to harm from (observed) impacts and (projected) risks and can be economic or
non- economic. (Mechler et al., 2018).
Maladaptive actions (Maladaptation): Actions that may lead to increased risk of adverse climate-related
outcomes, including via increased greenhouse gas (GHG) emissions, increased vulnerability to climate change, or
diminished welfare, now or in the future. Maladaptation is usually an unintended consequence.
Malnutrition: Deficiencies, excesses, or imbalances in a person’s intake of energy and/or nutrients. The term
malnutrition addresses three broad groups of conditions: undernutrition, which includes wasting (low weight-for-
height), stunting (low height-for-age) and underweight (low weight-for-age); micronutrient-related malnutrition,
which includes micronutrient deficiencies (a lack of important vitamins and minerals) or micronutrient excess; and
overweight, obesity and diet-related noncommunicable diseases (such as heart disease, stroke, diabetes and some
cancers) (WHO, 2018). Micronutrient deficiencies are sometimes termed ‘hidden hunger’ to emphasise that people
can be malnourished in the sense of deficient without being deficient in calories. Hidden hunger can apply even
where people are obese.
Marine heatwave: A period of extreme warm near-sea surface temperature that persists for days to months and can
extend up to thousands of kilometres.
Mean sea level: The surface level of the ocean at a particular point averaged over an extended period of time such
as a month or year. Mean sea level is often used as a national datum to which heights on land are referred.
Measurement: Processes of data collection over time, providing basic datasets, including associated accuracy and
precision, for the range of relevant variables. Possible data sources are field measurements, field observations,
detection through remote sensing and interviews (UN-REDD, 2009).
Megacity: Urban agglomerations with 10 million inhabitants or more. (United Nations, Department of Economic
and Social Affairs, Population Division (2019).
Megadrought: See Drought
Mental health: The state of well-being in which an individual realizes his or her own abilities, can cope with the
normal stresses of life, can work productively and is able to make a contribution to his or her community. (WHO,
XXXX)
Methane (CH4): One of the six greenhouse gases (GHGs) to be mitigated under the Kyoto Protocol. Methane is
the major component of natural gas and associated with all hydrocarbon fuels. Significant anthropogenic emissions
also occur as a result of animal husbandry and paddy rice production. Methane is also produced naturally where
organic matter decays under anaerobic conditions, such as in wetlands. Under future global warming, there is risk
of increased methane emissions from thawing permafrost, coastal wetlands and sub-sea gas hydrates.
Metric: A consistent measurement of a characteristic of an object or activity that is otherwise difficult to quantify.
Within the context of the evaluation of climate models, this is a quantitative measure of agreement between a
simulated and observed quantity which can be used to assess the performance of individual models.
Microclimate: Local climate at or near the Earth's surface.
Migrant: Any person who is moving or has moved across an international border or within a State away from
his/her habitual place of residence, regardless of (1) the person’s legal status; (2) whether the movement is
voluntary or involuntary; (3) what the causes for the movement are; or (4) what the length of the stay is. (IOM,
2018).
Migration (of humans): Movement of a person or a group of persons, either across an international border, or
within a State. It is a population movement, encompassing any kind of movement of people, whatever its length,
composition and causes; it includes migration of refugees, displaced persons, economic migrants, and persons
moving for other purposes, including family reunification (IOM, 2018).
Mitigation (of climate change): A human intervention to reduce emissions or enhance the sinks of greenhouse
gases.

20
Mitigation measures: In climate policy, mitigation measures are technologies, processes or practices that
contribute to mitigation, for example renewable energy technologies, waste minimization processes, and public
transport commuting practices.
Mitigation option: A technology or practice that reduces greenhouse gas (GHG) emissions or enhances sinks.
Mitigation scenario: A plausible description of the future that describes how the (studied) system responds to the
implementation of mitigation policies and measures.
Model Ensemble: A group of parallel model simulations characterising historical climate conditions, climate
predictions, or climate projections. Variation of the results across the ensemble members may give an estimate of
modelling-based uncertainty. Ensembles made with the same model but different initial conditions only
characterise the uncertainty associated with internal climate variability, whereas multi-model ensembles including
simulations by several models also include the impact of model differences. Perturbed parameter ensembles, in
which model parameters are varied in a systematic manner, aim to assess the uncertainty resulting from internal
model specifications within a single model. Remaining sources of uncertainty unaddressed with model ensembles
are related to systematic model errors or biases, which may be assessed from systematic comparisons of model
simulations with observations wherever available.
Monitoring and evaluation (M&E): Mechanisms put in place at national to local scales to respectively monitor
and evaluate efforts to reduce greenhouse gas emissions and/or adapt to the impacts of climate change with the aim
of systematically identifying, characterizing and assessing progress over time.
Monsoon: A monsoon is a tropical and subtropical seasonal reversal in both the surface winds and associated
precipitation, caused by differential heating between a continental-scale land mass and the adjacent ocean.
Monsoon rains occur mainly over land in summer.
Motivation (of an individual): An individual’s reason or reasons for acting in a particular way; individuals may
consider various consequences of actions, including financial, social, affective, and environmental consequences.
Motivation can arise from factors external or internal to the individual.
Mountains: A mountain is a landform formed through plate tectonics that rises above its surrounding area,
characterized by verticality and ruggedness such as gentle or steep sloping sides, sharp or rounded ridges, and a
high point called a peak or a summit. Mountain regions consist of mountains and mountain ranges as defined by
ruggedness, intermontane valleys, plateaus and tablelands, and hills and hilly forelands, together forming a complex
terrain. To delineate mountain regions a combination of terrain characteristics is used, such as elevation above sea
level, steepness of slope and relative relief or local elevational range. Various mountain characterizations using
different combinations of the above criteria applied to digital elevation models have been developed to arrive at
mountain area statistics, described and analysed in detail in Sayre et al, (2018), including characterizations termed
K1 (Kapos et al., 2000), K2 (Körner et al., 2011) and K3 (Karagulle et al., 2017).
Multilevel governance: See Governance
Narrative: Qualitative descriptions of plausible future world evolutions, describing the characteristics, general
logic and developments underlying a particular quantitative set of scenarios. Narratives are also referred to in the
literature as ‘storylines’.
Nationally Determined Contributions (NDCs): A term used under the United Nations Framework Convention on
Climate Change (UNFCCC) whereby a country that has joined the Paris Agreement outlines its plans for reducing
its emissions. Some countries’ NDCs also address how they will adapt to climate change impacts, and what support
they need from, or will provide to, other countries to adopt low-carbon pathways and to build climate resilience.
According to Article 4 paragraph 2 of the Paris Agreement, each Party shall prepare, communicate and maintain
successive NDCs that it intends to achieve.
Natural systems: The dynamic physical and biological components of the environment that would operate in the
absence of human impacts. Most, if not all, natural systems are also now affected by human activities to some
degree.

21
Nature-based solution (NBS): Actions to protect, sustainably manage and restore natural or modified ecosystems
that address societal challenges effectively and adaptively, simultaneously providing human well-being and
biodiversity benefits. (IUCN, 2016)
Net Primary Production (NPP): See Primary Production
New Urban Agenda: The New Urban Agenda was adopted at the United Nations Conference on Housing and
Sustainable Urban Development (Habitat III) in Quito, Ecuador, on 20 October 2016. It was endorsed by the United
Nations General Assembly at its sixty-eighth plenary meeting of the seventy-first session on 23 December 2016.
Non-climatic driver (Non-climate driver): An agent or process outside the climate system that influences a
human or natural system.
Non-communicable diseases: Non-communicable diseases (NCDs), also known as chronic diseases, tend to be of
long duration and are the result of a combination of genetic, physiological, environmental and behaviours factors.
The main types of NCDs are cardiovascular diseases (like heart attacks and stroke), cancers, chronic respiratory
diseases (such as chronic obstructive pulmonary disease and asthma) and diabetes (WHO)
Ocean : The interconnected body of saline water that covers 71% of the Earth's surface, contains 97% of the Earth's
water and provides 99% of the Earth's biologically-habitable space. It includes the Arctic, Atlantic, Indian, Pacific
and Southern Oceans, as well as their marginal seas and coastal waters.
Ocean acidification (OA): A reduction in the pH of the ocean, accompanied by other chemical changes (primarily
in the levels of carbonate and bicarbonate ions), over an extended period, typically decades or longer, which is
caused primarily by uptake of carbon dioxide (CO2) from the atmosphere, but can also be caused by other chemical
additions or subtractions from the ocean. Anthropogenic OA refers to the component of pH reduction that is caused
by human activity (IPCC, 2011, p. 37).
Ocean deoxygenation: The loss of oxygen in the ocean. It results from ocean warming, which reduces oxygen
solubility and increases oxygen consumption and stratification, thereby reducing the mixing of oxygen into the
ocean interior. Deoxygenation can also be exacerbated by the addition of excess nutrients in the coastal zone.
Ocean stratification ; See stratification
Oxygen Minimum Zone (OMZ): The midwater layer (200–1000 m) in the open ocean in which oxygen saturation
is the lowest in the ocean. The degree of oxygen depletion depends on the largely bacterial consumption of organic
matter and the distribution of the OMZs is influenced by large-scale ocean circulation. In coastal oceans, OMZs
extend to the shelves and may also affect benthic ecosystems.
Ozone (O3) The triatomic form of oxygen, and a gaseous atmospheric constituent. In the troposphere, O3 is
created both naturally and by photochemical reactions involving gases resulting from human activities (e.g., smog).
Tropospheric O3 acts as a greenhouse gas (GHG). In the stratosphere, O3 is created by the interaction between
solar ultraviolet radiation and molecular oxygen (O2). Stratospheric O3 plays a dominant role in the stratospheric
radiative balance. Its concentration is highest in the ozone layer.
Particulate matter (PM): Very small solid particles emitted during the combustion of biomass and fossil fuels.
PM may consist of a wide variety of substances. Of greatest concern for health are particulates of diameter less than
or equal to 10 nanometers, usually designated as PM10.
Particulates: Very small solid particles emitted during the combustion of fossil and biomass fuels. Particulates
may consist of a wide variety of substances. Of greatest concern for health are particulates of diameter less than or
equal to 10 nm, usually designated as PM10.
Pasture: Area covered with grass or other plants used or suitable for grazing of livestock; grassland.
Path dependence AR5-WG2 and WG3 The generic situation where decisions, events, or outcomes at one point in
time constrain adaptation, mitigation, or other actions or options at a later point in time.
Pathways: The temporal evolution of natural and/or human systems towards a future state. Pathway concepts range
from sets of quantitative and qualitative scenarios or narratives of potential futures to solution-oriented decision-
making processes to achieve desirable societal goals. Pathway approaches typically focus on biophysical, techno-

22
economic, and/or socio-behavioral trajectories and involve various dynamics, goals, and actors across different
scales.
Adaptation pathways: A series of adaptation choices involving trade-offs between short-term and long-term goals
and values. These are processes of deliberation to identify solutions that are meaningful to people in the context of
their daily lives and to avoid potential maladaptation.
Climate-resilient development pathways (CRDPs): Trajectories that strengthen sustainable development and
efforts to eradicate poverty and reduce inequalities while promoting fair and cross-scalar adaptation to and
resilience in a changing climate. They raise the ethics, equity, and feasibility aspects of the deep societal
transformation needed to drastically reduce emissions to limit global warming (e.g., to well below 2°C) and achieve
desirable and liveable futures and wellbeing for all.
Climate-resilient pathways: Iterative processes for managing change within complex systems in order to reduce
disruptions and enhance opportunities associated with climate change.
Development pathways: Trajectories based on an array of social, economic, cultural, technological, institutional,
and biophysical features that characterise the interactions between human and natural systems and outline visions
for the future, at a particular scale.
Emission pathways: Modelled trajectories of global anthropogenic emissions over the 21st century are termed
emission pathways.
Representative Concentration Pathways (RCPs): Scenarios that include time series of emissions and
concentrations of the full suite of greenhouse gases (GHGs) and aerosols and chemically active gases, as well as
land use/land cover (Moss et al., 2008; van Vuuren et al., 2011), The word representative signifies that each RCP
provides only one of many possible scenarios that would lead to the specific radiative forcing characteristics. The
term pathway emphasises the fact that not only the long-term concentration levels, but also the trajectory taken over
time to reach that outcome are of interest (Moss et al., 2010; van Vuuren et al., 2011). RCPs usually refer to the
portion of the concentration pathway extending up to 2100, for which integrated assessment models produced
corresponding emission scenarios. Extended concentration pathways describe extensions of the RCPs from 2100 to
2300 that were calculated using simple rules generated by stakeholder consultations, and do not represent fully
consistent scenarios. Four RCPs produced from Integrated assessment models were selected from the published
literature and are used in the Fifth IPCC Assessment and also used in this Assessment for comparison, spanning the
range from approximately below 2°C warming to high (>4°C) warming best-estimates by the end of the 21st
century: RCP2.6, RCP4.5 and RCP6.0 and RCP8.5.
• RCP2.6: One pathway where radiative forcing peaks at approximately 3 W m-2 and then declines to be
limited at 2.6 W m-2 in 2100 (the corresponding Extended Concentration Pathway, or ECP, has constant
emissions after 2100).
• RCP4.5 and RCP6.0: Two intermediate stabilisation pathways in which radiative forcing is limited at
approximately 4.5 W m-2 and 6.0 W m-2 in 2100 (the corresponding ECPs have constant concentrations
after 2150).
• RCP8.5: One high pathway which leads to >8.5 W m-2 in 2100 (the corresponding ECP has constant
emissions after 2100 until 2150 and constant concentrations after 2250).
Shared socio-economic pathways (SSPs) Shared socio-economic pathways (SSPs) have been developed to
complement the Representative concentration pathways (RCPs). By design, the RCP emission and concentration
pathways were stripped of their association with a certain socio-economic development. Different levels of
emissions and climate change along the dimension of the RCPs can hence be explored against the backdrop if
different socio-economic development pathways (SSPs) on the other dimension in a matrix. This integrative SSP-
RCP framework is now widely used in the climate impact and policy analysis literature (see e.g. http://iconics-
ssp.org), where climate projections obtained under the RCP scenarios are analysed against the backdrop of various
SSPs.
As several emission updates were due, a new set of emission scenarios was developed in conjunction with the
SSPs. Hence, the abbreviation SSP is now used for two things: On the one hand SSP1, SSP2,

23
SSP5 is used to denote the five socio-economic scenario families. On the other hand, the abbreviations SSP1-1.9,
SSP1-2.6, …, SSP5-8.5 are used to denote the newly developed emission scenarios that are the result of an SSP
implementation within an integrated assessment model. Those SSP scenarios are bare of climate policy assumption,
but in combination with so-called shared policy assumptions (SPAs), various approximate radiative forcing levels
of 1.9, 2.6, …, or 8.5 W m-2 are reached by the end of the century, respectively.
Sustainable development pathways (SDPs): Trajectories aimed at attaining the Sustainable Development Goals
(SDGs) in the short term and the goals of sustainable development in the long term. In the context of climate
change, such pathways denote trajectories that address social, environmental, and economic dimensions of
sustainable development, adaptation and mitigation, and transformation, in a generic sense or from a particular
methodological perspective such as integrated assessment models and scenario simulations.
Peat: Soft, porous or compressed, sedentary deposit of which a substantial portion is partly decomposed plant
material with high water content in the natural state (up to about 90 percent) (IPCC, 2013).
Peatlands: Peatlands are wetland ecosystems where soils are dominated by peat. In peatlands net primary
production exceeds organic matter decomposition as a result of waterlogged conditions, which leads to the
accumulation of peat.
Pelagic: The pelagic zone consists of the entire water column of the open ocean. It is subdivided into the
'epipelagic zone' (<200 m, the uppermost part of the ocean that receives enough sunlight to allow photosynthesis),
the 'mesopelagic zone' (200–1000 m depth) and the 'bathypelagic zone' (>1000 m depth). The term ‘pelagic’ can
also refer to organisms that live in the pelagic zone.
Pelagos : Organisms large and small living in the pelagic zones. Includes plankton (small) and nekton (free
swimming, large). See Benthos.
Percentiles: A partition value in a population distribution that a given percentage of the data values are below or
equal to. The 50th percentile corresponds to the median of the population. Percentiles are often used to estimate the
extremes of a distribution. For example, the 90th (10th) percentile may be used to refer to the threshold for the
upper (lower) extremes.
Peri-urban areas: Parts of a city that appear to be quite rural but are in reality strongly linked functionally to the
city in its daily activities.
Permafrost: Ground (soil or rock, and included ice and organic material) that remains at or below 0°C for at least
two consecutive years (Harris et al., 1988). Note that permafrost is defined via temperature rather than ice content
and, in some instances, may be ice-free.
Permafrost degradation: Decrease in the thickness and/or areal extent of permafrost.
Permafrost thaw: Progressive loss of ground ice in permafrost, usually due to input of heat. Thaw can occur over
decades to centuries over the entire depth of permafrost ground, with impacts occurring while thaw progresses.
During thaw, temperature fluctuations are subdued because energy is transferred by phase change between ice and
water. After the transition from permafrost to non-permafrost, ground can be described as thawed.
pH : A dimensionless measure of the acidity of a solution given by its concentration of hydrogen ions (H+). pH is
measured on a logarithmic scale where pH = -log10(H+). Thus, a pH decrease of 1 unit corresponds to a 10- fold
increase in the concentration of H+, or acidity.
Phenology: The relationship between biological phenomena that recur periodically (e.g., development stages,
migration) especially related to climate and seasonal changes.
Photosynthesis: The production of carbohydrates in plants, algae and some bacteria using the energy of light.
Carbon dioxide (CO2) is used as the carbon source.
Plankton: Microorganisms living in the upper layers of aquatic systems. A distinction is made between
phytoplankton, which depend on photosynthesis for their energy supply, and zooplankton, which feed on
phytoplankton.

24
Planned relocation (of humans): A form of human mobility response in the face of sea level rise and related
impacts. Planned relocation is typically initiated, supervised and implemented from national to local level and
involves small communities and individual assets but may also involve large populations. Also termed resettlement,
managed retreat, or managed realignment.
Plasticity (biology): Change in organismal trait values in response to an environmental cue, and which does not
require change in underlying DNA sequence.
Political economy: The set of interlinked relationships between people, the state, society and markets as defined by
law, politics, economics, customs and power that determine the outcome of trade and transactions and the
distribution of wealth in a country or economy.
Polycentric governance: See Governance
Potential Evapotranspiration: The potential rate of water loss without any limits imposed by the water supply.
Poverty: A complex concept with several definitions stemming from different schools of thought. It can refer to
material circumstances (such as need, pattern of deprivation or limited resources), economic conditions (such as
standard of living, inequality or economic position) and/or social relationships (such as social class, dependency,
exclusion, lack of basic security or lack of entitlement).
Poverty trap: Poverty trap is understood differently across disciplines. In the social sciences, the concept,
primarily employed at the individual, household, or community level, describes a situation in which escaping
poverty becomes impossible due to unproductive or inflexible resources. A poverty trap can also be seen as a
critical minimum asset threshold, below which families are unable to successfully educate their children, build up
their productive assets, and get out of poverty. Extreme poverty is itself a poverty trap, since poor persons lack the
means to participate meaningfully in society. In economics, the term poverty trap is often used at national scales,
referring to a self-perpetuating condition where an economy, caught in a vicious cycle, suffers from persistent
underdevelopment (Matsuyama, 2008). Many proposed models of poverty traps are found in the literature.
Pre-industrial (period): The multi-century period prior to the onset of large-scale industrial activity around 1750.
The reference period 1850–1900 is used to approximate pre-industrial global mean surface temperature (GMST).
Precursors: Atmospheric compounds that are not greenhouse gases (GHGs) or aerosols, but that have an effect on
GHG or aerosol concentrations by taking part in physical or chemical processes regulating their production or
destruction rates.
Predictability: The extent to which future states of a system may be predicted based on knowledge of current and
past states of the system. Because knowledge of the climate system's past and current states is generally imperfect,
as are the models that utilize this knowledge to produce a climate prediction, and because the climate system is
inherently nonlinear and chaotic, predictability of the climate system is inherently limited. Even with arbitrarily
accurate models and observations, there may still be limits to the predictability of such a nonlinear system (AMS,
2000).
Primary production: The synthesis of organic compounds by plants and microbes, on land or in the ocean,
primarily by photosynthesis using light and carbon dioxide (CO2) as sources of energy and carbon respectively. It
can also occur through chemosynthesis, using chemical energy, e.g., in deep sea vents.
Net Primary production (NPP): The difference between how much CO2 vegetation takes in during
photosynthesis (gross primary production) minus how much CO2 the plants release during respiration (IPBES,
2019. Global Assessment).
Procedural justice: See Justice
Projection: A potential future evolution of a quantity or set of quantities, often computed with the aid of a model.
Unlike predictions, projections are conditional on assumptions concerning, for example, future socio-economic and
technological developments that may or may not be realised.
Proxy: A proxy climate indicator is a record that is interpreted, using physical and biophysical principles, to
represent some combination of climate-related variations back in time. Climate-related data derived in this way are
referred to as proxy data. Examples of proxies include pollen analysis, tree ring records, speleothems,

25
characteristics of corals, and various data derived from marine sediments and ice cores. Proxy data can be
calibrated to provide quantitative climate information.
Public-Private Partnerships: Arrangements typified by joint working between the public and private sector. In the
broadest sense, they cover all types of collaboration across the interface between the public and private sectors to
deliver services or infrastructure.
Radiative forcing: The change in the net, downward minus upward, radiative flux (expressed in W m-2) at the
tropopause or top of atmosphere due to a change in an [external] driver of climate change, such as a change in the
concentration of carbon dioxide (CO2), the concentration of volcanic aerosols or in the output of the Sun. The
traditional radiative forcing is computed with all tropospheric properties held fixed at their unperturbed values, and
after allowing for stratospheric temperatures, if perturbed, to readjust to radiative-dynamical equilibrium. Radiative
forcing is called instantaneous if no change in stratospheric temperature is accounted for. The radiative forcing once
rapid adjustments are accounted for is termed the effective radiative forcing. Radiative forcing is not to be confused
with cloud radiative forcing, which describes an unrelated measure of the impact of clouds on the radiative flux at
the top of the atmosphere.
Reasons for Concern (RFCs): Elements of a classification framework, first developed in the IPCC Third
Assessment Report, which aims to facilitate judgments about what level of climate change may be dangerous (in
the language of Article 2 of the UNFCCC) by aggregating risks from various sectors, considering hazards,
exposures, vulnerabilities, capacities to adapt, and the resulting impacts.
Reference scenario: The scenario used as starting or reference point for a comparison between two or more
scenarios.
[Note 1: In many types of climate change research, reference scenarios reflect specific assumptions about patterns
of socio-economic development and may represent futures that assume no climate policies or specified climate
policies, for example those in place or planned at the time a study is carried out. Reference scenarios may also
represent futures with limited or no climate impacts or adaptation, to serve as a point of comparison for futures with
impacts and adaptation. These are also referred to as baseline scenarios in the literature.
Note 2: Reference scenarios can also be climate policy or impact scenarios, which in that case are taken as a point
of comparison to explore the implications of other features, e.g., of delay, technological options, policy design and
strategy or to explore the effects of additional impacts and adaptation beyond those represented in the reference
scenario.
Note 3: The term business as usual scenario has been used to describe a scenario that assumes no additional policies
beyond those currently in place and that patterns of socio-economic development are consistent with recent trends.
The term is now used less frequently than in the past.
Note 4: In climate change attribution or impact attribution research reference scenarios may refer to counterfactual
historical scenarios assuming no anthropogenic greenhouse gas emissions (climate change attribution) or no climate
change (impact attribution)]

Reforestation: Conversion to forest of land that has previously contained forests but that has been converted to
some other use.
[Note: For a discussion of the term forest and related terms such as afforestation, reforestation and deforestation in
the context of reporting and accounting Article 3.3 and 3.4 activities under the Kyoto Protocol, see 2013 Revised
Supplementary Methods and Good Practice Guidance Arising from the Kyoto Protocol.]--
Region: A relatively large-scale land or ocean area characterised by specific geographical and climatological
features. The climate of a land-based region is affected by regional and local scale features like topography, land
use characteristics and large water bodies, as well as remote influences from other regions, in addition to global
climate conditions. The IPCC defines a set of standard regions for analyses of observed climate trends and climate
model projections (see IPCC, 2018a, Figure 3.2; IPCC 2012a).

26
Regulation: A rule or order issued by governmental executive authorities or regulatory agencies and having the
force of law. Regulations implement policies and are mostly specific for particular groups of people, legal entities
or targeted activities. Regulation is also the act of designing and imposing rules or orders. Informational,
transactional, administrative and political constraints in practice limit the regulator's capability for implementing
preferred policies.
Relative humidity: The relative humidity specifies the ratio of actual water vapour pressure to that at saturation
with respect to liquid water or ice at the same temperature.
Renewable Energy (RE): Any form of energy from solar, geophysical or biological sources that is replenished by
natural processes at a rate that equals or exceeds its rate of use.
Reporting: The process of formal reporting of assessment results to the UNFCCC, according to predetermined
formats and according to established standards, especially the Intergovernmental Panel on Climate Change (IPCC)
Guidelines and GPG (Good Practice Guidance) (UN REDD, 2009).
Representative Concentration Pathways (RCPs): See Pathways
Reservoir: A component or components of the climate system where a greenhouse gas (GHG) or a precursor of a
greenhouse gas is stored (UNFCCC Article 1.7).
Residual risk: The risk that remains following adaptation and risk reduction efforts. See also Loss and Damage
Resilience: The capacity of interconnected social, economic and ecological systems to cope with a hazardous
event, trend or disturbance, responding or reorganising in ways that maintain their essential function, identity and
structure. Resilience is a positive attribute when it maintains capacity for adaptation, learning and/or transformation
(Arctic Council, 2016).
See also Hazard, Risk, and Vulnerability.
Resolution : In climate models, this term refers to the physical distance (metres or degrees) between each point on
the grid used to compute the equations. Temporal resolution refers to the time step or time elapsed between each
model computation of the equations.
Respiration: The process whereby living organisms convert organic matter to carbon dioxide (CO2), releasing
energy and consuming molecular oxygen.
Restoration: In environmental context, restoration involves human interventions to assist the recovery of an
ecosystem that has been previously degraded, damaged or destroyed.
Return period: An estimate of the average time interval between occurrences of an event (e.g., flood or extreme
rainfall) of (or below/above) a defined size or intensity.
Risk : The potential for adverse consequences for human or ecological systems, recognising 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. Relevant adverse consequences include those on
lives, livelihoods, health and wellbeing, economic, social and cultural assets and investments, infrastructure,
services (including ecosystem services), ecosystems and species.
In the context of climate change impacts, risks result from dynamic interactions between climate-related hazards
with the exposure and vulnerability of the affected human or ecological system to the hazards. Hazards, exposure
and vulnerability may each be subject to uncertainty in terms of magnitude and likelihood of occurrence, and each
may change over time and space due to socio-economic changes and human decision-making.
In the context of climate change responses, risks result from the potential for such responses not achieving the
intended objective(s), or from potential trade-offs with, or negative side-effects on, other societal objectives, such
as the Sustainable Development Goals (SDGs). Risks can arise for example from uncertainty in implementation,
effectiveness or outcomes of climate policy, climate-related investments, technology development or adoption, and
system transitions.
Risk assessment : The qualitative and/or quantitative scientific estimation of risks.

27
Risk management ; Plans, actions, strategies or policies to reduce the likelihood and/or magnitude of adverse
potential consequences, based on assessed or perceived risks.
Risk perception: The subjective judgment that people make about the characteristics and severity of a risk.
Risk transfer: The process of formally or informally shifting the financial consequences of particular risks from
one party to another whereby a household, community, enterprise, or state authority will obtain resources from the
other party after a disaster occurs, in exchange for ongoing or compensatory social or financial benefits provided to
that other party.
Runoff: The flow of water over the surface or through the subsurface, which typically originates from the part of
liquid precipitation and/or snow/ice melt that does not evaporate, transpire or refreeze, and returns to water bodies.
Salt-water intrusion/encroachment: Displacement of fresh surface water or groundwater by the advance of salt
water due to its greater density. This usually occurs in coastal and estuarine areas due to decreasing land-based
influence (e.g., from reduced runoff or groundwater recharge, or from excessive water withdrawals from aquifers)
or increasing marine influence (e.g., relative sea level rise).
Scenario: A plausible description of how the future may develop based on a coherent and internally consistent set
of assumptions about key driving forces (e.g., rate of technological change (TC), prices) and relationships. Note
that scenarios are neither predictions nor forecasts, but are used to provide a view of the implications of
developments and actions in a ‘what-if’ kind of investigation. In a broader sense, the term ‘scenarios’ is often used
to encompass ‘pathways’.
In the Sixth Assessment Report a minimum set of five scenarios is chosen to assist cross-working group
comparisons: SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP3-7.0 and SSP5-8.5. These are called illustrative marker SSP
scenarios and span a wide range of plausible futures from potentially below 1.5°C best-estimate warming to very
high warming in excess of 4°C over the course of this century. [See AR6 WGI Chapter 1, Cross-Chapter Box 1.5.
See also Pathways.]
Concentration scenario: A plausible representation of the future development of atmospheric concentrations of
substances that are radiatively active (e.g., greenhouse gases (GHGs), aerosols, tropospheric ozone), plus human-
induced land cover changes that can be radiatively active via albedo changes, and often used as input to a climate
model to compute climate projections.
Climate scenario: A plausible and often simplified representation of the future climate, based on an internally
consistent set of climatological relationships that has been constructed for explicit use in investigating the potential
consequences of anthropogenic climate change, often serving as input to impact models. Climate projections often
serve as the raw material for constructing climate scenarios, but climate scenarios usually require additional
information such as the observed current climate.
Emission scenario: A plausible representation of the future development of emissions of substances that are
radiativelyactive (e.g., greenhouse gases (GHGs), or aerosols) based on a coherent and internally consistent set of
assumptions about driving forces (such as demographic and socio-economic development, technological change,
energy and land use) and their key relationships. Concentration scenarios, derived from emission scenarios, are
often used as input to a climate model to compute climate projections.
Sea ice: Ice found at the sea surface that has originated from the freezing of seawater. Sea ice may be discontinuous
pieces (ice floes) moved on the ocean surface by wind and currents (pack ice), or a motionless sheet attached to the
coast (land-fast ice). Sea ice concentration is the fraction of the ocean covered by ice. Sea ice less than one year old
is called first-year ice. Perennial ice is sea ice that survives at least one summer. It may be subdivided into second-
year ice and multi-year ice, where multiyear ice has survived at least two summers.
Sea level change (sea level rise/sea level fall): Change to the height of sea level, both globally and locally (relative
sea level change) [at seasonal, annual,or longer time scales] due to (1) a change in ocean volume as a result of a
change in the mass of water in the ocean [(e.g., due to melt of glaciers and ice sheets)], (2) changes in ocean
volume as a result of changes in ocean water density [(e.g., expansion under warmer conditions)], (3) changes in
the shape of the ocean basins and changes in the Earth’s gravitational and rotational fields, and (4) local subsidence
or uplift of the land. Global mean sea level change resulting from change in the mass of the ocean is called

28
barystatic. The amount of barystatic sea level change due to the addition or removal of a mass of water is called its
sea level equivalent (SLE). Sea level changes, both globally and locally, resulting from changes in water density are
called steric. Density changes induced by temperature changes only are called thermosteric, while density changes
induced by salinity changes are called halosteric. Barystatic and steric sea level changes do not include the effect of
changes in the shape of ocean basins induced by the change in the ocean mass and its distribution.
Sea surface temperature (SST) : The subsurface bulk temperature in the top few metres of the ocean, measured
by ships, buoys, and drifters. From ships, measurements of water samples in buckets were mostly switched in the
1940s to samples from engine intake water. Satellite measurements of skin temperature (uppermost layer; a fraction
of a millimetre thick) in the infrared or the top centimeter or so in the microwave are also used, but must be
adjusted to be compatible with the bulk temperature.
Semi-arid zone: Areas where vegetation growth is constrained by limited water availability, often with short
growing seasons and high interannual variation in primary production. Annual precipitation ranges from 300 to 800
mm, depending on the occurrence of summer and winter rains.
Sendai Framework for Disaster Risk Reduction: The Sendai Framework for Disaster Risk Reduction 2015–2030
outlines seven clear targets and four priorities for action to prevent new, and to reduce existing, disaster risks. The
voluntary, non-binding agreement recognises that the State has the primary role to reduce disaster risk but that
responsibility should be shared with other stakeholders, including local government and the private sector. Its aim
is to achieve ‘substantial reduction of disaster risk and losses in lives, livelihoods and health and in the economic,
physical, social, cultural and environmental assets of persons, businesses, communities and countries.’
Sensitivity: The degree to which a system or species is affected, either adversely or beneficially, by climate
variability or change. The effect may be direct (e.g., a change in crop yield in response to a change in the mean,
range, or variability of temperature) or indirect (e.g., damages caused by an increase in the frequency of coastal
flooding due to sea level rise).
Sequestration: The process of storing carbon in a carbon pool.
Shared Socio-economic Pathways (SSPs) : See Pathways
Shelf seas : Relatively shallow water covering the shelf of continents or around islands. The limit of shelf seas is
conventionally considered as 200 m water depth at the continental shelf edge, where there is usually a steep slope to
the deep ocean floor. During glacial periods, most shelf seas are lost since they become land as the build-up of ice
sheets caused a decrease of global sea level.
Sink: Any process, activity or mechanism which removes a greenhouse gas, an aerosol or a precursor of a
greenhouse gas from the atmosphere (UNFCCC Article 1.8).
Small island developing states (SIDS): Small island developing states (SIDS), as recognized by the United
Nations OHRLLS (Office of the High Representative for the Least Developed Countries, Landlocked Developing
Countries and Small Island Developing States), are a distinct group of developing countries facing specific social,
economic and environmental vulnerabilities (UN-OHRLLS, 2011). They were recognized as a special case both for
their environment and development at the Rio Earth Summit in Brazil in 1992. Fifty-eight countries and territories
are presently classified as SIDS by the UN OHRLLS, with 38 being UN member states and 20 being Non- UN
Members or Associate Members of the Regional Commissions (UN-OHRLLS, 2018).
Snow cover extent: The areal extent of snow covered ground.
Snow water equivalent (SWE): The depth of liquid water that would result if a mass of snow melted completely.
Social inclusion: A process of improving the terms of participation in society, particularly for people who are
disadvantaged, through enhancing opportunities, access to resources, and respect for rights (UN, DESA 2016).
Social justice: See Justice
Social learning: A process of social interaction through which people learn new behaviours, capacities, values, and
attitudes.

29
Social protection: In the context of development aid and climate policy, social protection usually describes public
and private initiatives that provide income or consumption transfers to the poor, protect the vulnerable against
livelihood risks, and enhance the social status and rights of the marginalized, with the overall objective of reducing
the economic and social vulnerability of poor, vulnerable, and marginalized groups (Devereux and Sabates-
Wheeler, 2004). In other contexts, social protection may be used synonymously with social policy and can be
described as all public and private initiatives that provide access to services, such as health, education, or housing,
or income and consumption transfers to people. Social protection policies protect the poor and vulnerable against
livelihood risks and enhance the social status and rights of the marginalized, as well as prevent vulnerable people
from falling into poverty.
Social-ecological system: An integrated system that includes human societies and ecosystems, in which humans
are part of nature. The functions of such a system arise from the interactions and interdependence of the social and
ecological subsystems. The system’s structure is characterised by reciprocal feedbacks, emphasising that humans
must be seen as a part of, not apart from, nature (Arctic Council, 2016; Berkes and Folke, 1998).
Societal (social) transformation : See Transformation
Socio-economic scenario: A scenario that describes a possible future in terms of population, gross domestic
product (GDP), and other socio-economic factors relevant to understanding the implications of climate change.
Socio-technical transitions : Where technological change is associated with social systems and the two are
inextricably linked.
Soil erosion: The displacement of the soil by the action of water or wind. Soil erosion is a major process of land
degradation.
Soil moisture: Water stored in the soil in liquid or frozen form. Root-zone soil moisture is of most relevance for
plant activity.
Soil organic carbon: Carbon contained in soil organic matter.
Soil organic matter: The organic component of soil, comprising plant and animal residue at various stages of
decomposition, and soil organisms.
Solar Radiation Modification (SRM): Solar radiation modification (SRM) refers to a range of radiation
modification measures not related to greenhouse gas (GHG) mitigation that seek to limit global warming. Most
methods involve reducing the amount of incoming solar radiation reaching the surface, but others also act on the
longwave radiation budget by reducing optical thickness and cloud lifetime. (IPCC 2018a)
Solution space: The set of biophysical, cultural, socio-economic, and political-institutional dimensions within
which opportunities and constraints determine why, how, when, and who acts to reduce climate risks. Within these
dimensions, there are ‘hard’ (unsurpassable) limits and ‘soft’ (surpassable) limits. The boundaries of the solution
space are path dependent, contested, and in constant flux (Haasnoot et. al. 2020).
Source: Any process or activity which releases a greenhouse gas, an aerosol or a precursor of a greenhouse gas into
the atmosphere (UNFCCC Article 1.9).
Southern Ocean: The ocean region encircling Antarctica that connects the Atlantic, Indian and Pacific Oceans
together, allowing inter-ocean exchange. This region is the main source of much of the deep water of the world’s
ocean and also provides the primary return pathway for this deep water to the surface (Marshall and Speer, 2012;
Toggweiler and Samuels, 1995). The drawing up of deep waters and the subsequent transport into the ocean interior
has major consequences for the global heat, nutrient, and carbon balances, as well as the Antarctic cryosphere and
marine ecosystems.
Spatial and temporal scales: Climate may vary on a large range of spatial and temporal scales. Spatial scales may
range from local (less than 100 000 km2), through regional (100 000 to 10 million km2) to continental (10 to 100
million km2). Temporal scales may range from seasonal to geological (up to hundreds of millions of years).
Standards: Set of rules or codes mandating or defining product performance (e.g., grades, dimensions,
characteristics, test methods, and rules for use). Product, technology or performance standards establish minimum

30
requirements for affected products or technologies. Standards impose reductions in greenhouse gas (GHG)
emissions associated with the manufacture or use of the products and/or application of the technology.
Storm surge: The temporary increase, at a particular locality, in the height of the sea due to extreme
meteorological conditions (low atmospheric pressure and/or strong winds). The storm surge is defined as being the
excess above the level expected from the tidal variation alone at that time and place.
Storyline: A way of making sense of a situation or a series of events through the construction of a set of
explanatory elements. Usually it is built on logical or causal reasoning. In climate research, the term storyline is
used both in connection to scenarios as related to a future trajectory of the climate and human systems or to a
weather or climate event. In this context, storylines can be used to describe plural, conditional possible futures or
explanations of a current situation, in contrast to single, definitive futures or explanations.
Physical climate storyline: A self-consistent and plausible unfolding of a physical trajectory of the climate system,
or a weather or climate event, on timescales from hours to multiple decades (Shepherd et al., 2018). Through this,
storylines explore, illustrate and communicate uncertainties in the climate system response to forcing and in
internal variability.
Scenario storyline: A narrative description of a scenario (or family of scenarios), highlighting the main scenario
characteristics, relationships between key driving forces and the dynamics of their evolution.
Stranded assets: Assets exposed to devaluations or conversion to ‘liabilities’ because of unanticipated changes in
their initially expected revenues due to innovations and/or evolutions of the business context, including changes in
public regulations at the domestic and international levels.
Stratification: Process of forming of layers of (ocean) water with different properties such as salinity, density and
temperature that act as barrier for water mixing. The strengthening of near-surface stratification generally results in
warmer surface waters, decreased oxygen levels in deeper water, and intensification of ocean acidification (OA) in
the upper ocean.
Streamflow : Water flow within a river channel, for example, expressed in m3 s-1. A synonym for river discharge.
Stressors: Events and trends, often not climate-related, that have an important effect on the system exposed and
can increase vulnerability to climate-related risk.
Sustainability: Involves ensuring the persistence of natural and human systems, implying the continuous
functioning of ecosystems, the conservation of high biodiversity, the recycling of natural resources and, in the
human sector, successful application of justice and equity.
Sustainable development (SD): Development that meets the needs of the present without compromising the ability
of future generations to meet their own needs (WCED, 1987) and balances social, economic and environmental
concerns.
Sustainable Development Goals (SDGs) : The 17 global goals for development for all countries established by the
United Nations through a participatory process and elaborated in the 2030 Agenda for Sustainable Development,
including ending poverty and hunger; ensuring health and wellbeing, education, gender equality, clean water and
energy, and decent work; building and ensuring resilient and sustainable infrastructure, cities and consumption;
reducing inequalities; protecting land and water ecosystems; promoting peace, justice and partnerships; and taking
urgent action on climate change.
Sustainable development pathways (SDPs): See Pathways
Sustainable land management: The stewardship and use of land resources, including soils, water, animals and
plants, to meet changing human needs, while simultaneously ensuring the long-term productive potential of these
resources and the maintenance of their environmental functions (Adapted from WOCAT, undated). Sympagic
organisms and habitats related to the sea ice, analogous to ‘pelagic’ (> water-column’) or ‘benthic’ (> ‘sea- floor’).
Teleconnection: A statistical association between climate variables at widely separated, geographically-fixed
spatial locations. Teleconnections are caused by large spatial structures such as basin-wide coupled modes of
ocean-atmosphere variability, Ross by wave-trains, mid-latitude jets, and storm tracks.

31
Temperature overshoot: The temporary exceedance of a specified level of global warming, such as 1.5°C.
Overshoot implies a peak followed by a decline in global warming, achieved through anthropogenic removal of
carbon dioxide (CO2) exceeding remaining CO2 emissions globally.
Tier: In the context of the IPCC Guidelines for National Greenhouse Gas Inventories, a tier represents a level of
methodological complexity. Usually three tiers are provided. Tier 1 is the basic method, Tier 2 intermediate and
Tier 3 most demanding in terms of complexity and data requirements. Tiers 2 and 3 are sometimes referred to as
higher tier methods and are generally considered to be more accurate (IPCC, 2019).
Tipping point: A level of change in system properties beyond which a system reorganises, often [in a non-linear
manner/abruptly], and does not return to the initial state even if the drivers of the change are abated. For the climate
system, the term refers to a critical threshold [at/beyond] which global or regional climate changes from one stable
state to another stable state. Tipping points are also used when referring to impact: the term can imply that an
impact tipping point is (about to be) reached in a natural or human system.
Transformation: A change in the fundamental attributes of natural and human systems. Societal (social)
transformation. A profound and often deliberate shift initiated by communities toward sustainability, facilitated by
changes in individual and collective values and behaviours, and a fairer balance of political, cultural, and
institutional power in society.
Transformational adaptation: See Adaptation
Transformative change: A system-wide change that requires more than technological change through
consideration of social and economic factors that, with technology, can bring about rapid change at scale.
Transition: The process of changing from one state or condition to another in a given period of time. Transition
can occur in individuals, firms, cities, regions and nations, and can be based on incremental or transformative
change.
Tree line: The upper limit of tree growth in mountains or at high latitudes. It is more elevated or more poleward
than the forest line.
Trend (statistical): In this report, the word trend designates a change, generally monotonic in time, in the value of
a variable.
Tropical cyclone: The general term for a strong, cyclonic-scale disturbance that originates over tropical oceans.
Distinguished from weaker systems (often named tropical disturbances or depressions) by exceeding a threshold
wind speed. A tropical storm is a tropical cyclone with one-minute average surface winds between 18 and 32 m s-
1. Beyond 32 m s-1, a tropical cyclone is called a hurricane, typhoon, or cyclone, depending on geographic
location.
Tsunami: A wave, or train of waves, produced by a disturbance such as a submarine earthquake displacing the sea
floor, a landslide, a volcanic eruption, or an asteroid impact.
Tundra: A treeless biome characteristic of polar and alpine regions.
Uncertainty: A state of incomplete knowledge that can result from a lack of information or from disagreement
about what is known or even knowable. It may have many types of sources, from imprecision in the data to
ambiguously defined concepts or terminology, incomplete understanding of critical processes, or uncertain
projections of human behaviour. Uncertainty can therefore be represented by quantitative measures (e.g., a
probability density function) or by qualitative statements (e.g., reflecting the judgment of a team of experts) (see
Moss and Schneider, 2000; IPCC, 2004; Mastrandrea et al., 2010).
United Nations Framework Convention on Climate Change (UNFCCC): The UNFCCC was adopted in May
1992 and opened for signature at the 1992 Earth Summit in Rio de Janeiro. It entered into force in March 1994 and
as of May 2018 had 197 Parties (196 States and the European Union). The Convention’s ultimate objective is the
‘stabilisation of greenhouse gas concentrations in the atmosphere at a level that would prevent dangerous
anthropogenic interference with the climate system’. The provisions of the Convention are pursued and
implemented by two treaties: the Kyoto Protocol and the Paris Agreement.

32
Uptake: The transfer of substances (such as carbon) or energy (e.g., heat) from one compartment of a system to
another; for example, in the Earth system from the atmosphere to the ocean or to the land.
Upwelling region: A region of an ocean where cold, typically nutrient-rich waters well up from the deep ocean.
Urban and Peri-urban agriculture: The cultivation of crops and rearing of animals for food and other uses within
and surrounding the boundaries of cities, including fisheries and forestry (EPRS, 2014).
Urban heat island (UHI):The relative warmth of a city compared with surrounding rural areas, associated with
changes in runoff,effects on heat retention, and changes in surface albedo.
Values and Beliefs: Fundamental attitudes about what is important, good, and right; strongly held principles or
qualities intrinsically valuable or desirable, often enshrined in laws, traditions, and religions. Examples include
human rights, subsistence, and equitable distribution of costs and benefits of climate policies (Hulme, 2009, 2018;
Nakashima et al., 2012; UNFCCC, 1992; UN Universal Declaration of Human Rights, 1948).
Vector-borne disease: Illnesses caused by parasites, viruses and bacteria that are transmitted by various vectors
(e.g. mosquitoes, sandflies, triatomine bugs, blackflies, ticks, tsetse flies, mites, snails and lice) (UNEP 2018).
Ventilation (ocean): The exchange of ocean properties with the atmospheric surface layer such that property
concentrations are brought closer to equilibrium values with the atmosphere (AMS, 2000), and the processes that
propagate these properties into the ocean interior.
Vulnerability: The propensity or predisposition to be adversely affected. Vulnerability encompasses a variety of
concepts and elements including sensitivity or susceptibility to harm and lack of capacity to cope and adapt.
Vulnerability index: A metric characterizing the vulnerability of a system. A climate vulnerability index is
typically derived by combining, with or without weighting, several indicators assumed to represent vulnerability.
Warm days/warm nights: Days where maximum temperature, or nights where minimum temperature, exceeds the
90th percentile, where the respective temperature distributions are generally defined with respect to the 1961-1990
reference period.
Water-borne disease: Illnesses that transmitted through contact with or consumption of unsafe or contaminated
water (UNEP 2018).
Water security: The capacity of a population to safeguard sustainable access to adequate quantities of acceptable
quality water for sustaining livelihoods, human well-being, and socio-economic development, for ensuring
protection against water-borne pollution and water-related disasters, and for preserving ecosystems in a climate of
peace and political stability (UN-Water, 2013).
Water-use efficiency: Carbon gain by photosynthesis per unit of water lost by evapotranspiration. It can be
expressed on a short- term basis as the ratio of photosynthetic carbon gain per unit transpirational water loss, or on
a seasonal basis as the ratio of net primary production or agricultural yield to the amount of water used.
Weathering: The gradual removal of atmospheric CO2 through dissolution of silicate and carbonate rocks.
Weathering may involve physical processes (mechanical weathering) or chemical activity (chemical weathering).
Well-being: A state of existence that fulfils various human needs, including material living conditions and quality
of life, as well as the ability to pursue one’s goals, to thrive, and feel satisfied with one’s life. Ecosystem well-being
refers to the ability of ecosystems to maintain their diversity and quality.
Wetland: Land that is covered or saturated by water for all or part of the year (e.g., peatland).

33
United Nations A/RES/70/1

Distr.: General
General Assembly 21 October 2015

Seventieth session
Agenda items 15 and 116

Resolution adopted by the General Assembly on 25 September 2015


[without reference to a Main Committee (A/70/L.1)]

70/1. Transforming our world: the 2030 Agenda for


Sustainable Development

The General Assembly


Adopts the following outcome document of the United Nations summit for the
adoption of the post-2015 development agenda:

Transforming our world: the 2030 Agenda for Sustainable


Development

Preamble
This Agenda is a plan of action for people, planet and prosperity. It also seeks
to strengthen universal peace in larger freedom. We recognize that eradicating
poverty in all its forms and dimensions, including extreme poverty, is the greatest
global challenge and an indispensable requirement for sustainable development.
All countries and all stakeholders, acting in collaborative partnership, will
implement this plan. We are resolved to free the human race from the tyranny of
poverty and want and to heal and secure our planet. We are determined to take the
bold and transformative steps which are urgently needed to shift the world on to a
sustainable and resilient path. As we embark on this collective journey, we pledge
that no one will be left behind.
The 17 Sustainable Development Goals and 169 targets which we are
announcing today demonstrate the scale and ambition of this new universal Agenda.
They seek to build on the Millennium Development Goals and complete what they
did not achieve. They seek to realize the human rights of all and to achieve gender
equality and the empowerment of all women and girls. They are integrated and
indivisible and balance the three dimensions of sustainable development: the
economic, social and environmental.
The Goals and targets will stimulate action over the next 15 years in areas of
critical importance for humanity and the planet.

15-16301 (E)
*1516301* Please recycle

34
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

People

We are determined to end poverty and hunger, in all their forms and
dimensions, and to ensure that all human beings can fulfil their potential in dignity
and equality and in a healthy environment.

Planet

We are determined to protect the planet from degradation, including through


sustainable consumption and production, sustainably managing its natural resources
and taking urgent action on climate change, so that it can support the needs of the
present and future generations.

Prosperity

We are determined to ensure that all human beings can enjoy prosperous and
fulfilling lives and that economic, social and technological progress occurs in
harmony with nature.

Peace

We are determined to foster peaceful, just and inclusive societies which are
free from fear and violence. There can be no sustainable development without peace
and no peace without sustainable development.

Partnership

We are determined to mobilize the means required to implement this Agenda


through a revitalized Global Partnership for Sustainable Development, based on a
spirit of strengthened global solidarity, focused in particular on the needs of the
poorest and most vulnerable and with the participation of all countries, all
stakeholders and all people.

The interlinkages and integrated nature of the Sustainable Development Goals


are of crucial importance in ensuring that the purpose of the new Agenda is realized.
If we realize our ambitions across the full extent of the Agenda, the lives of all will
be profoundly improved and our world will be transformed for the better.

2/35

35
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

Sustainable Development Goals

Goal 1. End poverty in all its forms everywhere

Goal 2. End hunger, achieve food security and improved nutrition and
promote sustainable agriculture

Goal 3. Ensure healthy lives and promote well-being for all at all ages

Goal 4. Ensure inclusive and equitable quality education and promote


lifelong learning opportunities for all

Goal 5. Achieve gender equality and empower all women and girls

Goal 6. Ensure availability and sustainable management of water and


sanitation for all

Goal 7 Ensure access to affordable, reliable, sustainable and modern


energy for all

Goal 8. Promote sustained, inclusive and sustainable economic growth,


full and productive employment and decent work for all

Goal 9. Build resilient infrastructure, promote inclusive and sustainable


industrialization and foster innovation

Goal 10. Reduce inequality within and among countries

Goal 11. Make cities and human settlements inclusive, safe, resilient and
sustainable

Goal 12. Ensure sustainable consumption and production patterns

Goal 13. Take urgent action to combat climate change and its impacts*

Goal 14. Conserve and sustainably use the oceans, seas and marine
resources for sustainable development

Goal 15. Protect, restore and promote sustainable use of terrestrial


ecosystems, sustainably manage forests, combat desertification,
and halt and reverse land degradation and halt biodiversity loss

Goal 16. Promote peaceful and inclusive societies for sustainable


development, provide access to justice for all and build
effective, accountable and inclusive institutions at all levels

Goal 17. Strengthen the means of implementation and revitalize the


Global Partnership for Sustainable Development

*
Acknowledging that the United Nations Framework Convention on Climate Change
is the primary international, intergovernmental forum for negotiating the global
response to climate change .

14/35

36
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

Goal 1. End poverty in all its forms everywhere


1.1 By 2030, eradicate extreme poverty for all people everywhere, currently
measured as people living on less than $1.25 a day
1.2 By 2030, reduce at least by half the proportion of men, women and children of
all ages living in poverty in all its dimensions according to national definitions
1.3 Implement nationally appropriate social protection systems and measures for
all, including floors, and by 2030 achieve substantial coverage of the poor and the
vulnerable
1.4 By 2030, ensure that all men and women, in particular the poor and the
vulnerable, have equal rights to economic resources, as well as access to basic
services, ownership and control over land and other forms of property, inheritance,
natural resources, appropriate new technology and financial services, including
microfinance
1.5 By 2030, build the resilience of the poor and those in vulnerable situations and
reduce their exposure and vulnerability to climate-related extreme events and other
economic, social and environmental shocks and disasters

1.a Ensure significant mobilization of resources from a variety of sources,


including through enhanced development cooperation, in order to provide adequate
and predictable means for developing countries, in particular least developed
countries, to implement programmes and policies to end poverty in all its
dimensions
1.b Create sound policy frameworks at the national, regional and international
levels, based on pro-poor and gender-sensitive development strategies, to support
accelerated investment in poverty eradication actions

Goal 2. End hunger, achieve food security and improved nutrition and
promote sustainable agriculture
2.1 By 2030, end hunger and ensure access by all people, in particular the poor
and people in vulnerable situations, including infants, to safe, nutritious and
sufficient food all year round
2.2 By 2030, end all forms of malnutrition, including achieving, by 2025, the
internationally agreed targets on stunting and wasting in children under 5 years of
age, and address the nutritional needs of adolescent girls, pregnant and lactating
women and older persons
2.3 By 2030, double the agricultural productivity and incomes of small-scale food
producers, in particular women, indigenous peoples, family farmers, pastoralists and
fishers, including through secure and equal access to land, other productive
resources and inputs, knowledge, financial services, markets and opportunities for
value addition and non-farm employment
2.4 By 2030, ensure sustainable food production systems and implement resilient
agricultural practices that increase productivity and production, that help maintain
ecosystems, that strengthen capacity for adaptation to climate change, extreme
weather, drought, flooding and other disasters and that progressively improve land
and soil quality
2.5 By 2020, maintain the genetic diversity of seeds, cultivated plants and farmed
and domesticated animals and their related wild species, including through soundly

15/35

37
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

managed and diversified seed and plant banks at the national, regional and
international levels, and promote access to and fair and equitable sharing of benefits
arising from the utilization of genetic resources and associated traditional
knowledge, as internationally agreed

2.a Increase investment, including through enhanced international cooperation, in


rural infrastructure, agricultural research and extension services, technology
development and plant and livestock gene banks in order to enhance agricultural
productive capacity in developing countries, in particular least developed countries
2.b Correct and prevent trade restrictions and distortions in world agricultural
markets, including through the parallel elimination of all forms of agricultural
export subsidies and all export measures with equivalent effect, in accordance with
the mandate of the Doha Development Round
2.c Adopt measures to ensure the proper functioning of food commodity markets
and their derivatives and facilitate timely access to market information, including on
food reserves, in order to help limit extreme food price volatility

Goal 3. Ensure healthy lives and promote well-being for all at all ages
3.1 By 2030, reduce the global maternal mortality ratio to less than 70 per 100,000
live births
3.2 By 2030, end preventable deaths of newborns and children under 5 years of
age, with all countries aiming to reduce neonatal mortality to at least as low as
12 per 1,000 live births and under-5 mortality to at least as low as 25 per 1,000 live
births
3.3 By 2030, end the epidemics of AIDS, tuberculosis, malaria and neglected
tropical diseases and combat hepatitis, water-borne diseases and other
communicable diseases
3.4 By 2030, reduce by one third premature mortality from non-communicable
diseases through prevention and treatment and promote mental health and well-
being
3.5 Strengthen the prevention and treatment of substance abuse, including narcotic
drug abuse and harmful use of alcohol
3.6 By 2020, halve the number of global deaths and injuries from road traffic
accidents
3.7 By 2030, ensure universal access to sexual and reproductive health-care
services, including for family planning, information and education, and the
integration of reproductive health into national strategies and programmes
3.8 Achieve universal health coverage, including financial risk protection, access
to quality essential health-care services and access to safe, effective, quality and
affordable essential medicines and vaccines for all
3.9 By 2030, substantially reduce the number of deaths and illnesses from
hazardous chemicals and air, water and soil pollution and contamination

3.a Strengthen the implementation of the World Health Organization Framework


Convention on Tobacco Control in all countries, as appropriate
3.b Support the research and development of vaccines and medicines for the
communicable and non-communicable diseases that primarily affect developing

16/35

38
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

countries, provide access to affordable essential medicines and vaccines, in


accordance with the Doha Declaration on the TRIPS Agreement and Public Health,
which affirms the right of developing countries to use to the full the provisions in
the Agreement on Trade-Related Aspects of Intellectual Property Rights regarding
flexibilities to protect public health, and, in particular, provide access to medicines
for all
3.c Substantially increase health financing and the recruitment, development,
training and retention of the health workforce in developing countries, especially in
least developed countries and small island developing States
3.d Strengthen the capacity of all countries, in particular developing countries, for
early warning, risk reduction and management of national and global health risks

Goal 4. Ensure inclusive and equitable quality education and promote lifelong
learning opportunities for all
4.1 By 2030, ensure that all girls and boys complete free, equitable and quality
primary and secondary education leading to relevant and effective learning outcomes
4.2 By 2030, ensure that all girls and boys have access to quality early childhood
development, care and pre-primary education so that they are ready for primary
education
4.3 By 2030, ensure equal access for all women and men to affordable and quality
technical, vocational and tertiary education, including university
4.4 By 2030, substantially increase the number of youth and adults who have
relevant skills, including technical and vocational skills, for employment, decent
jobs and entrepreneurship
4.5 By 2030, eliminate gender disparities in education and ensure equal access to
all levels of education and vocational training for the vulnerable, including persons
with disabilities, indigenous peoples and children in vulnerable situations
4.6 By 2030, ensure that all youth and a substantial proportion of adults, both men
and women, achieve literacy and numeracy
4.7 By 2030, ensure that all learners acquire the knowledge and skills needed to
promote sustainable development, including, among others, through education for
sustainable development and sustainable lifestyles, human rights, gender equality,
promotion of a culture of peace and non-violence, global citizenship and appreciation
of cultural diversity and of culture’s contribution to sustainable development

4.a Build and upgrade education facilities that are child, disability and gender
sensitive and provide safe, non-violent, inclusive and effective learning
environments for all
4.b By 2020, substantially expand globally the number of scholarships available to
developing countries, in particular least developed countries, small island
developing States and African countries, for enrolment in higher education,
including vocational training and information and communications technology,
technical, engineering and scientific programmes, in developed countries and other
developing countries
4.c By 2030, substantially increase the supply of qualified teachers, including
through international cooperation for teacher training in developing countries,
especially least developed countries and small island developing States

17/35

39
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

Goal 5. Achieve gender equality and empower all women and girls
5.1 End all forms of discrimination against all women and girls everywhere
5.2 Eliminate all forms of violence against all women and girls in the public and
private spheres, including trafficking and sexual and other types of exploitation
5.3 Eliminate all harmful practices, such as child, early and forced marriage and
female genital mutilation
5.4 Recognize and value unpaid care and domestic work through the provision of
public services, infrastructure and social protection policies and the promotion of
shared responsibility within the household and the family as nationally appropriate
5.5 Ensure women’s full and effective participation and equal opportunities for
leadership at all levels of decision-making in political, economic and public life
5.6 Ensure universal access to sexual and reproductive health and reproductive
rights as agreed in accordance with the Programme of Action of the International
Conference on Population and Development and the Beijing Platform for Action and
the outcome documents of their review conferences

5.a Undertake reforms to give women equal rights to economic resources, as well
as access to ownership and control over land and other forms of property, financial
services, inheritance and natural resources, in accordance with national laws
5.b Enhance the use of enabling technology, in particular information and
communications technology, to promote the empowerment of women
5.c Adopt and strengthen sound policies and enforceable legislation for the
promotion of gender equality and the empowerment of all women and girls at all
levels

Goal 6. Ensure availability and sustainable management of water and


sanitation for all
6.1 By 2030, achieve universal and equitable access to safe and affordable
drinking water for all
6.2 By 2030, achieve access to adequate and equitable sanitation and hygiene for
all and end open defecation, paying special attention to the needs of women and
girls and those in vulnerable situations
6.3 By 2030, improve water quality by reducing pollution, eliminating dumping
and minimizing release of hazardous chemicals and materials, halving the
proportion of untreated wastewater and substantially increasing recycling and safe
reuse globally
6.4 By 2030, substantially increase water-use efficiency across all sectors and
ensure sustainable withdrawals and supply of freshwater to address water scarcity
and substantially reduce the number of people suffering from water scarcity
6.5 By 2030, implement integrated water resources management at all levels,
including through transboundary cooperation as appropriate
6.6 By 2020, protect and restore water-related ecosystems, including mountains,
forests, wetlands, rivers, aquifers and lakes

18/35

40
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

6.a By 2030, expand international cooperation and capacity-building support to


developing countries in water- and sanitation-related activities and programmes,
including water harvesting, desalination, water efficiency, wastewater treatment,
recycling and reuse technologies
6.b Support and strengthen the participation of local communities in improving
water and sanitation management

Goal 7. Ensure access to affordable, reliable, sustainable and modern energy


for all
7.1 By 2030, ensure universal access to affordable, reliable and modern energy
services
7.2 By 2030, increase substantially the share of renewable energy in the global
energy mix
7.3 By 2030, double the global rate of improvement in energy efficiency

7.a By 2030, enhance international cooperation to facilitate access to clean energy


research and technology, including renewable energy, energy efficiency and
advanced and cleaner fossil-fuel technology, and promote investment in energy
infrastructure and clean energy technology
7.b By 2030, expand infrastructure and upgrade technology for supplying modern
and sustainable energy services for all in developing countries, in particular least
developed countries, small island developing States and landlocked developing
countries, in accordance with their respective programmes of support

Goal 8. Promote sustained, inclusive and sustainable economic growth, full


and productive employment and decent work for all
8.1 Sustain per capita economic growth in accordance with national circumstances
and, in particular, at least 7 per cent gross domestic product growth per annum in
the least developed countries
8.2 Achieve higher levels of economic productivity through diversification,
technological upgrading and innovation, including through a focus on high-value
added and labour-intensive sectors
8.3 Promote development-oriented policies that support productive activities,
decent job creation, entrepreneurship, creativity and innovation, and encourage the
formalization and growth of micro-, small- and medium-sized enterprises, including
through access to financial services
8.4 Improve progressively, through 2030, global resource efficiency in
consumption and production and endeavour to decouple economic growth from
environmental degradation, in accordance with the 10-Year Framework of
Programmes on Sustainable Consumption and Production, with developed countries
taking the lead
8.5 By 2030, achieve full and productive employment and decent work for all
women and men, including for young people and persons with disabilities, and
equal pay for work of equal value
8.6 By 2020, substantially reduce the proportion of youth not in employment,
education or training

19/35

41
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

8.7 Take immediate and effective measures to eradicate forced labour, end modern
slavery and human trafficking and secure the prohibition and elimination of the
worst forms of child labour, including recruitment and use of child soldiers, and by
2025 end child labour in all its forms
8.8 Protect labour rights and promote safe and secure working environments for
all workers, including migrant workers, in particular women migrants, and those in
precarious employment
8.9 By 2030, devise and implement policies to promote sustainable tourism that
creates jobs and promotes local culture and products
8.10 Strengthen the capacity of domestic financial institutions to encourage and
expand access to banking, insurance and financial services for all

8.a Increase Aid for Trade support for developing countries, in particular least
developed countries, including through the Enhanced Integrated Framework for
Trade-related Technical Assistance to Least Developed Countries
8.b By 2020, develop and operationalize a global strategy for youth employment
and implement the Global Jobs Pact of the International Labour Organization

Goal 9. Build resilient infrastructure, promote inclusive and sustainable


industrialization and foster innovation
9.1 Develop quality, reliable, sustainable and resilient infrastructure, including
regional and transborder infrastructure, to support economic development and
human well-being, with a focus on affordable and equitable access for all
9.2 Promote inclusive and sustainable industrialization and, by 2030, significantly
raise industry’s share of employment and gross domestic product, in line with
national circumstances, and double its share in least developed countries
9.3 Increase the access of small-scale industrial and other enterprises, in particular
in developing countries, to financial services, including affordable credit, and their
integration into value chains and markets
9.4 By 2030, upgrade infrastructure and retrofit industries to make them
sustainable, with increased resource-use efficiency and greater adoption of clean
and environmentally sound technologies and industrial processes, with all countries
taking action in accordance with their respective capabilities
9.5 Enhance scientific research, upgrade the technological capabilities of
industrial sectors in all countries, in particular developing countries, including, by
2030, encouraging innovation and substantially increasing the number of research
and development workers per 1 million people and public and private research and
development spending

9.a Facilitate sustainable and resilient infrastructure development in developing


countries through enhanced financial, technological and technical support to African
countries, least developed countries, landlocked developing countries and small
island developing States
9.b Support domestic technology development, research and innovation in
developing countries, including by ensuring a conducive policy environment for,
inter alia, industrial diversification and value addition to commodities

20/35

42
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

9.c Significantly increase access to information and communications technology


and strive to provide universal and affordable access to the Internet in least
developed countries by 2020

Goal 10. Reduce inequality within and among countries


10.1 By 2030, progressively achieve and sustain income growth of the bottom
40 per cent of the population at a rate higher than the national average
10.2 By 2030, empower and promote the social, economic and political inclusion of
all, irrespective of age, sex, disability, race, ethnicity, origin, religion or economic
or other status
10.3 Ensure equal opportunity and reduce inequalities of outcome, including by
eliminating discriminatory laws, policies and practices and promoting appropriate
legislation, policies and action in this regard
10.4 Adopt policies, especially fiscal, wage and social protection policies, and
progressively achieve greater equality
10.5 Improve the regulation and monitoring of global financial markets and
institutions and strengthen the implementation of such regulations
10.6 Ensure enhanced representation and voice for developing countries in
decision-making in global international economic and financial institutions in order
to deliver more effective, credible, accountable and legitimate institutions
10.7 Facilitate orderly, safe, regular and responsible migration and mobility of
people, including through the implementation of planned and well-managed
migration policies

10.a Implement the principle of special and differential treatment for developing
countries, in particular least developed countries, in accordance with World Trade
Organization agreements
10.b Encourage official development assistance and financial flows, including
foreign direct investment, to States where the need is greatest, in particular least
developed countries, African countries, small island developing States and
landlocked developing countries, in accordance with their national plans and
programmes
10.c By 2030, reduce to less than 3 per cent the transaction costs of migrant
remittances and eliminate remittance corridors with costs higher than 5 per cent

Goal 11. Make cities and human settlements inclusive, safe, resilient
and sustainable
11.1 By 2030, ensure access for all to adequate, safe and affordable housing and
basic services and upgrade slums
11.2 By 2030, provide access to safe, affordable, accessible and sustainable
transport systems for all, improving road safety, notably by expanding public
transport, with special attention to the needs of those in vulnerable situations,
women, children, persons with disabilities and older persons
11.3 By 2030, enhance inclusive and sustainable urbanization and capacity for
participatory, integrated and sustainable human settlement planning and
management in all countries

21/35

43
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

11.4 Strengthen efforts to protect and safeguard the world’s cultural and natural
heritage
11.5 By 2030, significantly reduce the number of deaths and the number of people
affected and substantially decrease the direct economic losses relative to global
gross domestic product caused by disasters, including water-related disasters, with a
focus on protecting the poor and people in vulnerable situations
11.6 By 2030, reduce the adverse per capita environmental impact of cities,
including by paying special attention to air quality and municipal and other waste
management
11.7 By 2030, provide universal access to safe, inclusive and accessible, green and
public spaces, in particular for women and children, older persons and persons with
disabilities

11.a Support positive economic, social and environmental links between urban,
peri-urban and rural areas by strengthening national and regional development
planning
11.b By 2020, substantially increase the number of cities and human settlements
adopting and implementing integrated policies and plans towards inclusion, resource
efficiency, mitigation and adaptation to climate change, resilience to disasters, and
develop and implement, in line with the Sendai Framework for Disaster Risk
Reduction 2015–2030, holistic disaster risk management at all levels
11.c Support least developed countries, including through financial and technical
assistance, in building sustainable and resilient buildings utilizing local materials

Goal 12. Ensure sustainable consumption and production patterns


12.1 Implement the 10-Year Framework of Programmes on Sustainable
Consumption and Production Patterns, all countries taking action, with developed
countries taking the lead, taking into account the development and capabilities of
developing countries
12.2 By 2030, achieve the sustainable management and efficient use of natural
resources
12.3 By 2030, halve per capita global food waste at the retail and consumer levels
and reduce food losses along production and supply chains, including post-harvest
losses
12.4 By 2020, achieve the environmentally sound management of chemicals and all
wastes throughout their life cycle, in accordance with agreed international
frameworks, and significantly reduce their release to air, water and soil in order to
minimize their adverse impacts on human health and the environment
12.5 By 2030, substantially reduce waste generation through prevention, reduction,
recycling and reuse
12.6 Encourage companies, especially large and transnational companies, to adopt
sustainable practices and to integrate sustainability information into their reporting
cycle
12.7 Promote public procurement practices that are sustainable, in accordance with
national policies and priorities

22/35

44
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

12.8 By 2030, ensure that people everywhere have the relevant information and
awareness for sustainable development and lifestyles in harmony with nature

12.a Support developing countries to strengthen their scientific and technological


capacity to move towards more sustainable patterns of consumption and production
12.b Develop and implement tools to monitor sustainable development impacts for
sustainable tourism that creates jobs and promotes local culture and products
12.c Rationalize inefficient fossil-fuel subsidies that encourage wasteful
consumption by removing market distortions, in accordance with national
circumstances, including by restructuring taxation and phasing out those harmful
subsidies, where they exist, to reflect their environmental impacts, taking fully into
account the specific needs and conditions of developing countries and minimizing
the possible adverse impacts on their development in a manner that protects the poor
and the affected communities

Goal 13. Take urgent action to combat climate change and its impacts *

13.1 Strengthen resilience and adaptive capacity to climate-related hazards and


natural disasters in all countries
13.2 Integrate climate change measures into national policies, strategies and
planning
13.3 Improve education, awareness-raising and human and institutional capacity on
climate change mitigation, adaptation, impact reduction and early warning

13.a Implement the commitment undertaken by developed-country parties to the


United Nations Framework Convention on Climate Change to a goal of mobilizing
jointly $100 billion annually by 2020 from all sources to address the needs of
developing countries in the context of meaningful mitigation actions and
transparency on implementation and fully operationalize the Green Climate Fund
through its capitalization as soon as possible
13.b Promote mechanisms for raising capacity for effective climate change-related
planning and management in least developed countries and small island developing
States, including focusing on women, youth and local and marginalized
communities

Goal 14. Conserve and sustainably use the oceans, seas and marine resources
for sustainable development
14.1 By 2025, prevent and significantly reduce marine pollution of all kinds, in
particular from land-based activities, including marine debris and nutrient pollution
14.2 By 2020, sustainably manage and protect marine and coastal ecosystems to
avoid significant adverse impacts, including by strengthening their resilience, and
take action for their restoration in order to achieve healthy and productive oceans
14.3 Minimize and address the impacts of ocean acidification, including through
enhanced scientific cooperation at all levels

*
Acknowledging that the United Nations Framework Convention on Climate Change is the
primary international, intergovernmental forum for negotiating the global response to climate
change.

23/35

45
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

14.4 By 2020, effectively regulate harvesting and end overfishing, illegal,


unreported and unregulated fishing and destructive fishing practices and implement
science-based management plans, in order to restore fish stocks in the shortest time
feasible, at least to levels that can produce maximum sustainable yield as
determined by their biological characteristics
14.5 By 2020, conserve at least 10 per cent of coastal and marine areas, consistent
with national and international law and based on the best available scientific
information
14.6 By 2020, prohibit certain forms of fisheries subsidies which contribute to
overcapacity and overfishing, eliminate subsidies that contribute to illegal,
unreported and unregulated fishing and refrain from introducing new such subsidies,
recognizing that appropriate and effective special and differential treatment for
developing and least developed countries should be an integral part of the World
Trade Organization fisheries subsidies negotiation 16
14.7 By 2030, increase the economic benefits to small island developing States and
least developed countries from the sustainable use of marine resources, including
through sustainable management of fisheries, aquaculture and tourism

14.a Increase scientific knowledge, develop research capacity and transfer marine
technology, taking into account the Intergovernmental Oceanographic Commission
Criteria and Guidelines on the Transfer of Marine Technology, in order to improve
ocean health and to enhance the contribution of marine biodiversity to the
development of developing countries, in particular small island developing States
and least developed countries
14.b Provide access for small-scale artisanal fishers to marine resources and markets
14.c Enhance the conservation and sustainable use of oceans and their resources by
implementing international law as reflected in the United Nations Convention on the
Law of the Sea, which provides the legal framework for the conservation and
sustainable use of oceans and their resources, as recalled in paragraph 158 of “The
future we want”

Goal 15. Protect, restore and promote sustainable use of terrestrial ecosystems,
sustainably manage forests, combat desertification, and halt and
reverse land degradation and halt biodiversity loss
15.1 By 2020, ensure the conservation, restoration and sustainable use of terrestrial
and inland freshwater ecosystems and their services, in particular forests, wetlands,
mountains and drylands, in line with obligations under international agreements
15.2 By 2020, promote the implementation of sustainable management of all types
of forests, halt deforestation, restore degraded forests and substantially increase
afforestation and reforestation globally
15.3 By 2030, combat desertification, restore degraded land and soil, including land
affected by desertification, drought and floods, and strive to achieve a land
degradation-neutral world

_______________
16
Taking into account ongoing World Trade Organization negotiations, the Doha Development Agenda
and the Hong Kong ministerial mandate.

24/35

46
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

15.4 By 2030, ensure the conservation of mountain ecosystems, including their


biodiversity, in order to enhance their capacity to provide benefits that are essential
for sustainable development
15.5 Take urgent and significant action to reduce the degradation of natural
habitats, halt the loss of biodiversity and, by 2020, protect and prevent the
extinction of threatened species
15.6 Promote fair and equitable sharing of the benefits arising from the utilization
of genetic resources and promote appropriate access to such resources, as
internationally agreed
15.7 Take urgent action to end poaching and trafficking of protected species of flora
and fauna and address both demand and supply of illegal wildlife products
15.8 By 2020, introduce measures to prevent the introduction and significantly
reduce the impact of invasive alien species on land and water ecosystems and
control or eradicate the priority species
15.9 By 2020, integrate ecosystem and biodiversity values into national and local
planning, development processes, poverty reduction strategies and accounts

15.a Mobilize and significantly increase financial resources from all sources to
conserve and sustainably use biodiversity and ecosystems
15.b Mobilize significant resources from all sources and at all levels to finance
sustainable forest management and provide adequate incentives to developing
countries to advance such management, including for conservation and reforestation
15.c Enhance global support for efforts to combat poaching and trafficking of
protected species, including by increasing the capacity of local communities to
pursue sustainable livelihood opportunities

Goal 16. Promote peaceful and inclusive societies for sustainable development,
provide access to justice for all and build effective, accountable and
inclusive institutions at all levels
16.1 Significantly reduce all forms of violence and related death rates everywhere
16.2 End abuse, exploitation, trafficking and all forms of violence against and
torture of children
16.3 Promote the rule of law at the national and international levels and ensure
equal access to justice for all
16.4 By 2030, significantly reduce illicit financial and arms flows, strengthen the
recovery and return of stolen assets and combat all forms of organized crime
16.5 Substantially reduce corruption and bribery in all their forms
16.6 Develop effective, accountable and transparent institutions at all levels
16.7 Ensure responsive, inclusive, participatory and representative decision-
making at all levels
16.8 Broaden and strengthen the participation of developing countries in the
institutions of global governance
16.9 By 2030, provide legal identity for all, including birth registration

25/35

47
A/RES/70/1 Transforming our world: the 2030 Agenda for Sustainable Development

16.10 Ensure public access to information and protect fundamental freedoms, in


accordance with national legislation and international agreements

16.a Strengthen relevant national institutions, including through international


cooperation, for building capacity at all levels, in particular in developing countries,
to prevent violence and combat terrorism and crime
16.b Promote and enforce non-discriminatory laws and policies for sustainable
development

Goal 17. Strengthen the means of implementation and revitalize the Global
Partnership for Sustainable Development
Finance
17.1 Strengthen domestic resource mobilization, including through international
support to developing countries, to improve domestic capacity for tax and other
revenue collection
17.2 Developed countries to implement fully their official development assistance
commitments, including the commitment by many developed countries to achieve
the target of 0.7 per cent of gross national income for official development
assistance (ODA/GNI) to developing countries and 0.15 to 0.20 per cent of
ODA/GNI to least developed countries; ODA providers are encouraged to consider
setting a target to provide at least 0.20 per cent of ODA/GNI to least developed
countries
17.3 Mobilize additional financial resources for developing countries from multiple
sources
17.4 Assist developing countries in attaining long-term debt sustainability through
coordinated policies aimed at fostering debt financing, debt relief and debt
restructuring, as appropriate, and address the external debt of highly indebted poor
countries to reduce debt distress
17.5 Adopt and implement investment promotion regimes for least developed
countries

Technology
17.6 Enhance North-South, South-South and triangular regional and international
cooperation on and access to science, technology and innovation and enhance
knowledge sharing on mutually agreed terms, including through improved
coordination among existing mechanisms, in particular at the United Nations level,
and through a global technology facilitation mechanism
17.7 Promote the development, transfer, dissemination and diffusion of
environmentally sound technologies to developing countries on favourable terms,
including on concessional and preferential terms, as mutually agreed
17.8 Fully operationalize the technology bank and science, technology and
innovation capacity-building mechanism for least developed countries by 2017 and
enhance the use of enabling technology, in particular information and
communications technology

26/35

48
Transforming our world: the 2030 Agenda for Sustainable Development A/RES/70/1

Capacity-building
17.9 Enhance international support for implementing effective and targeted
capacity-building in developing countries to support national plans to implement all
the Sustainable Development Goals, including through North-South, South-South
and triangular cooperation

Trade
17.10 Promote a universal, rules-based, open, non-discriminatory and equitable
multilateral trading system under the World Trade Organization, including through
the conclusion of negotiations under its Doha Development Agenda
17.11 Significantly increase the exports of developing countries, in particular with
a view to doubling the least developed countries’ share of global exports by 2020
17.12 Realize timely implementation of duty-free and quota-free market access on
a lasting basis for all least developed countries, consistent with World Trade
Organization decisions, including by ensuring that preferential rules of origin
applicable to imports from least developed countries are transparent and simple, and
contribute to facilitating market access

Systemic issues
Policy and institutional coherence
17.13 Enhance global macroeconomic stability, including through policy
coordination and policy coherence
17.14 Enhance policy coherence for sustainable development
17.15 Respect each country’s policy space and leadership to establish and
implement policies for poverty eradication and sustainable development

Multi-stakeholder partnerships
17.16 Enhance the Global Partnership for Sustainable Development, complemented
by multi-stakeholder partnerships that mobilize and share knowledge, expertise,
technology and financial resources, to support the achievement of the Sustainable
Development Goals in all countries, in particular developing countries
17.17 Encourage and promote effective public, public-private and civil society
partnerships, building on the experience and resourcing strategies of partnerships

Data, monitoring and accountability


17.18 By 2020, enhance capacity-building support to developing countries,
including for least developed countries and small island developing States, to
increase significantly the availability of high-quality, timely and reliable data
disaggregated by income, gender, age, race, ethnicity, migratory status, disability,
geographic location and other characteristics relevant in national contexts
17.19 By 2030, build on existing initiatives to develop measurements of progress
on sustainable development that complement gross domestic product, and support
statistical capacity-building in developing countries

27/35

49
03 Environmental Movements

1.0 Deep Ecology


Deep Ecology is a radical environmental philosophy and political movement founded upon the holistic belief
that all living things have an equal right to life, or have subjective or objective intrinsic values. It has two
defining philosophical ideals. First, self-realization, which emphasizes a broadening and deepening of the self
toward a sense of personal identity that allows each being's potential to be realized. Second, biological
egalitarianism, the principle that humans have no more right to live than any other organism—all living things
are equally valuable and deserve the same consideration. In accordance with these principles, deep ecology
is considered a biocentric (living-centered), rather than an anthropocentric (human-centered), worldview.
The term deep ecology was coined by Norwegian professor of philosophy and accomplished mountaineer
Arne Naess in his 1973 article “The Shallow and the Deep, Long-Range Ecology Movements: A Summary.”
Naess' earlier scholarly pursuits had focused on semantics, unraveling the logic of language and investigating
the theoretical reasoning behind positivist science. His later writings on environmental philosophy were
influenced heavily by the work of Baruch Spinoza, Aldo Leopold, Rachel Carson, and Mahatma Gandhi, as
well as Taoist, Buddhist, and Native American belief systems, particularly their emphasis on organic unity
and their rejection of reductive binaries.
Naess later expanded these foundational lessons into the eight points of the deep ecology platform:
1. The well-being and flourishing of human and nonhuman life on Earth have value in themselves.
These values are independent of the usefulness of the nonhuman world for human purposes.
2. Richness and diversity of life forms contribute to the realization of these values and are also values
in themselves.
3. Humans have no right to reduce this richness and diversity except to satisfy vital needs.
4. The flourishing of human life and cultures is compatible with a substantial decrease of the human
population. The flourishing of nonhuman life requires such a decrease.
5. Present human interference with the nonhuman world is excessive, and the situation is rapidly
worsening.
6. Policies must therefore be changed. These policies affect basic economic, technological, and
ideological structures. The resulting state of affairs will be deeply different from the present.
7. The ideological change is mainly that of appreciating life qualities (dwelling in situations of inherent
value) rather than adhering to an increasingly higher standard of living. There will be a profound
awareness of the difference between big and great.
8. Those who subscribe to the foregoing points have an obligation directly or indirectly to try to
implement the necessary changes.
Naess described his approach as “deep” because it was concerned with fundamental questions of
humanity's role as part of the biosphere and demanded changes to all facets of human life. He dismissed
other environmental worldviews— conservationists, ecosocialists, and environmental justice advocates—
as “shallow” for adhering to a dualistic worldview in which the concerns and priorities of humans and
nonhumans are considered separately. Deep ecologists Bill Devall and George Sessions contrasted the
shallow and deep worldviews according to the following criteria:

Dominant Worldview Deep Ecology


Dominance over nature Harmony with nature
Natural environment as resource for humans All nature has intrinsic worth
Economic growth for human populations Simple material needs
Belief in ample resource reserves Earth “supplies” are limited
High-tech progress and solutions Appropriate technology
Consumerism Doing with enough
Nation-states as governing units Bioregions as governing units

50
Deep Ecology1
Deep Ecology is defined [8] as:
• a philosophy based on our sacred relationship with Earth and all beings
• an international movement for a viable future
• a path for self realisation
• a compass for daily action

Deep Ecology Supports:


• continuing inquiry into the appropriate human roles on our planet
• root cause analysis of unsustainable practices
• reduction of human consumption
• conservation and restoration of ecosystems
• a life of committed action for Earth

The Platform Principles of the Deep Ecology Movement: The eight points

i. The well-being and flourishing of human and nonhuman Life on Earth have value in themselves
(synonyms: intrinsic value, inherent value). These values are independent of the usefulness of
the nonhuman world for human purposes.
ii. Richness and diversity of life forms contribute to the realizations of these values and are also
values in themselves.
iii. Humans have no right to reduce this richness and diversity except to satisfy vital human needs.
iv. The flourishing of human life and cultures is compatible with a substantial decrease of human
population. The flourishing of nonhuman life requires such a decrease.
v. Present human interference with the nonhuman world is excessive, and the situation is rapidly
worsening.
vi. Policies must therefore be changed. These policies affect basic economic, technological, and
ideological structures. The resulting state of affairs will be deeply different from the present.
vii. The ideological change is mainly that of appreciating life quality (dwelling in situations of inherent
value) rather than adhering to an increasingly higher standard of living. There will be a profound
awareness of the difference between big and great.
viii. Those who subscribe to the foregoing points have an obligation to directly or indirectly try to
implement the necessary changes.

Another person starting from their own ultimate premise would use the same logical techniques to define their
own platform, and vice versa. This a good method of highlighting weaknesses, inconsistencies and
assumptions in a personal belief system, or to go deeper and not accept or present "shallow" arguments.

Deep versus Shallow Ecology


A number of key terms and slogans from the environmental debate will clarify the contrast between the
shallow and the deep ecology movements.

A. Pollution
Shallow Approach: Technology seeks to purify the air and water and to spread pollution more evenly. Laws
limit permissible pollution. Polluting industries are preferably exported to developing countries.

Deep Approach: Pollution is evaluated from a biospheric point of view, not focusing exclusively on its effects
on human health, but rather on life as a whole, including the life conditions of every species and system. The
shallow reaction to acid rain, for example, is to tend to avoid action by demanding more research and the
attempt to find species of trees which will tolerate high acidity etc. The deep approach concentrates on what
is going on in the total ecosystem and calls for a high priority fight against the economic conditions and the
technology responsible for producing the acid rain. The long range concerns are one hundred years, at least.

1
Downloaded from http://www.webnb.btinternet.co.uk/deep.htm on May 18, 2017

51
The priority is to fight the deep causes of pollution, not merely the superficial, short-range effects. The Third
and Fourth World countries cannot afford to pay the total costs of the war against pollution in their regions;
consequently they require the assistance of the First and Second World Countries. Exporting pollution is not
only a crime against humanity, it is a crime against life in general.

B. Resources
Shallow Approach: The emphasis is upon resources for humans, especially for the present generation in
affluent societies. In this view, the resources of the earth belong to those who have the technology to exploit
them. There is confidence that resources will not be depleted because as they get rarer, a high market price
will conserve them and substitutes will be found through technological progress. Further plants, animals, and
natural objects are valuable only as resources for humans. If no human use is known or seems likely ever to
be found, it does not matter if they are destroyed.

Deep Approach: The concern here is with resources and habits for all life forms and for their own sake. No
natural object is conceived of solely as a resource. This leads, them to a critical evaluation of human modes
of production and consumption. The question arises to what extent does an increase in production and
consumption foster ultimate human values. To what extend does it satisfy ital needs locally or globally? How
can economic, legal, and educational institutions be changed to counteract destructive increase? How can
resource use serve the quality of life rather than the economic standard of living as generally promoted by
consumerism? Form a deep perspective, there is an emphasis upon an ecosystem approach rather than the
consideration merely of isolated life forms or local situations? There is a long range maximal perspective of
time and place.

C. Population
Shallow Approach: The threat of (human) “over population” is seen mainly as a problem for developing
countries. One condones or even applauds population increases in one’s own country for short-sighted
economic, military, or other reasons; an increase in the number of humans is considered as valuable in itself
or as economically profitably. The issue of an “optimum population” for humans is discussed without reference
to the question of an “optimum population” for other life forms. The destruction of wild habitats caused by
increasing human population is accepoted as an inevitable evil, and drastic decreases of wildlife forms tend
to be accepted insofar as species are not driven to extinction. Further, the social relations of animals are
ignored. A long term substantial reduction of the global human population is not seen to be a desirable goal.
In addition, the right is claimed to defend one’s borders against “illegal aliens” regardless of what the
population pressures are elsewhere.

Deep Approach: It is recognized that excessive pressures on planetary life stem from the human population
explosion. The pressure stemming from the industrial societies is a major factor, and population reduction
must have the highest priority in those societies.

D. Cultural Diversity and Appropriate Technology


Shallow Approach: Industrialization of the Western industrial type is held to be the goal of developing
countries. The universal adoption of Western technology is held to be compatible with cultural diversity,
together with the conservation of the positive elements (from Western perspective) of present non-industrial
societies. There is a low estimate of deep cultural differences in non-industrial societies which deviate
significantly from contemporary Western standards.

Deep Approach: Protection of non-industrial cultures from invasion of industrial societies. The goals of the
former should not be seen as promoting lifestyles similar to those in the rich countries. Deep cultural diversity
is an analogue on the human level to the biological richness and diversity of life-forms. A high priority should
be given to cultural anthropology in general education programs in industrial societies.

There should be limits on the impact of Western technology upon present existing non- industrial countries
and the Fourth World should be defended against foreign domination. Political and economic policies should
favour subcultures within industrial societies. Local, soft technologies should allow for a basic cultural
assessment of any technical innovations, together with freely expressed criticism of so-called advanced
technology when this has the potential to be culturally destructive.

E. Land and sea ethics


Shallow Approach: Landscapes, ecosystems, rivers and other whole entities of nature are conceptually cut
into fragments, thus disregarding larger units and comprehensive gestalts. These fragments are regarded as

52
the properties and resources of individuals, organizations or states. Conservation is argued in terms of
multiple use and cost/benefit analysis. The social costs and long term global ecological costs of resource
extraction and use are usually not considered. Wildlife management is conceived of as conserving nature for
future generations of humans. Soil erosion or the deterioration of ground water quality, for example, is noted
as a human loss, but a strong belief in future technological progress makes deep changes seem unnecessary.

Deep approach: The earth does not belong to humans. For example, the Norwegian landscapes, rivers, flora
and fauna and the neighboring sea are not property of Norwegians. Similarly, the oil under the North Sea or
anywhere else does not belong to any state or to humanity. And the free nature surrounding a local community
does not belong to the local community.

Humans only inhabit the lands, using resources to satisfy vital needs. And if their non-vital needs come in
conflict with the vital needs of nonhumans, then humans should defer to the latter. The ecological destruction
now going on will not be cured by a technological fix. Current arrogant notions in industrial (and other)
societies must be resisted.

F. Education and the scientific enterprise


Shallow Approach: The degradation of the environment and resource depletion requires the training of more
and more experts who can provide advice concerning how to continue combining economic growth with
maintaining a healthy environment. We are likely to need an increasingly more dominating and manipulative
technology to manage the plant when global economic growth makes further environmental degradation
inevitable. The scientific enterprise must continue giving priority to the hard sciences (physics and chemistry).
High educational standards with intense competition in the relevant tough areas of learning will be required.

Deep Approach: If sane ecological policies are adopted, then education should concentrate on an increased
sensitivity to non-consumptive goods and on such consumables where there is enough for all. Education
should therefore counteract the excessive emphasis upon things with a price tag. There should be a shift in
concentration from the hard to the soft sciences which stress the importance of the local and global cultures.
The educational objective of the World Conservation Strategy (building support for conservation) should be
given a high priority, but within the deeper framework of respect for the biosphere.

In the future, there will be no shallow environmental movement if deep policies are increasingly adopted by
governments and thus no need for a special deep ecological social movement.

Deep Experience

Deep experience, or it might be termed a moment of enlightenment, is often what gets a person started along
a deep ecological path.

Stephan Harding [10] cites Aldo Leopold, from his book A Sand County Almanac, as an example of this. For
Leopold, the experience was of sufficient intensity to trigger a total reorientation in his life's work as a wildlife
manager and ecologist. In the 1920s he had been appointed by the US government to develop a rational,
scientific policy for eradicating the wolf from the entire United States.

As a wildlife manager of those times, Leopold adhered to the unquestioning belief that humans were superior
to the rest of nature, and were thus morally justified in manipulating it as much as was required in order to
maximise human welfare.

"Thinking like a mountain" has become a key phrase in the deep ecology movement, as evidenced by the
book Thinking like a mountain – towards a council of all beings [2].

Also for Arne Naess a key influence has been his deep relationship to Hallingskarvet mountain in central
Norway, where, in 1937, he built a simple cabin at the place called Tvergastein (crossed stones). It is in this
place: high up, totally isolated, with commanding views of landscape down below, with Arctic storms
threatening to blow away his roof, that most of his important work in deep ecology has been done. In this
inhospitable retreat, under snow and ice for most of the year, where only lichen and tiny alpine flowers grow,
Arne Naess has spent a total of more than ten years, watching, climbing, thinking, writing, and adoring the
mountain. [10]

53
2.0 Eco-centric

The first Principle of the Deep Ecology Movement "The well-being and flourishing of human and nonhuman
Life on Earth have value in themselves" is expanded by Naess [12]:

"Rejection of the man-in-environment image in favor of the relational, oral-field image. Organisms as knots in
the biospherical net or field of intrinsic relations. An intrinsic relation between two things A and B is such that
the relation belongs to the definitions or basic constitutions of A and B. so that without the relation, A and B
are no longer the same things. The total-field model dissolves not only the man-in-environment concept, but
every compact thing in-milieu concept except when talking at a superficial or preliminary level of
communication"

Human culture and philosophy, especially in the currently dominant Western tradition, has focused on human
values and eschewed the natural world.

A distinctive aspect of the deep ecology movement is its recognition of the inherent value of all other living
beings, and of the inherent worth of diversity of all kinds. This awareness is used to shape environmental
policies and actions. Those who work for social changes based on this recognition are motivated by love of
Nature as well as for humans. They try to be caring in all their dealings. They recognise that we cannot go on
with industrial culture’s business as usual. We must make fundamental changes in basic values and practices
or we will destroy the diversity and beauty of the world, and its ability to support diverse human cultures.

The Deep Ecology movement is not misanthropic, see first principle, although some people have written
about it as such; nor does being affected more by the suffering of family or other humans than other creatures
mean that a person cannot be a supporter of the movement.

Ralph Metzner [4] points to two meanings of Anthropocentric. One is the literal meaning "human centred"
which it has been pointed out, that as humans, can be our only perspective on the world. Metzner says the
deep ecologists’ critique of anthropocentrism has an implicit meaning of assumed superiority and right to
dominate others, this can also be referred to as human chauvinism or speciesism.

3.0 Eco-feminism & Social Ecology

In an interview Michael E. Zimmerman, Professor of Philosophy at Tulane University, New Orleans [11] said
about eco-feminism:

"There are many ecofeminists - people like Joanna Macy for example - who would call themselves deep
ecologists, but there are some ecofeminists who've made an important claim against it. They say the real
problem isn't anthropocentrism but androcentrism - man- centeredness. They say that 10,000 years of
patriarchy is ultimately responsible for the destruction of the biosphere and the development of authoritarian
practices, both socially and environmentally."

"Deep ecologists concede that patriarchy has been responsible for a lot of violence against women and
nature. But while they oppose the oppression of women and promote egalitarian social relations, deep
ecologists also warn that getting rid of patriarchy would not necessarily cure the problem, because you can
imagine a society with fairly egalitarian social relationships where nature is still used instrumentally."

There are other social ecologists who see that the problem of the environmental crisis is directly linked to
authoritarianism and hierarchy. This includes issues like racism, sexism, third world exploitation, mistreatment
of other marginalised groups etc, as well as nature.

Naess commented on this issue [12] "Ecologically responsible policies are concerned only in part with
pollution and resource depletion. There are deeper concerns which touch upon principles of diversity,
complexity, autonomy, decentralization, symbiosis, egalitarianism, and classlessness."

Deep ecology, perhaps by trying to avoid anthropocentrism, puts Nature, or Gaia, before human society, but
that is not to deny the value of humans and their culture, and the need to change current behaviour. Naess
said [12] "Diversity of human ways of life is in part due to (intended or unintended) exploitation and
suppression on the part of certain groups. The exploiter lives differently from the exploited, but both are
adversely affected in their potentialities of self-realization. The principle of diversity does not cover differences

54
due merely to certain attitudes or behaviors forcibly blocked or restrained. The principles of ecological
egalitarianism and of symbiosis support the same anticlass posture. The ecological attitude favors the
extension of all three principles to any group conflicts; including those of today between developing and
developed nations. The three principles also favor extreme caution toward any overall plans for the future,
except those consistent with wide and widening classless diversity."

The 4th principle of the Deep Ecology Movement is "The flourishing of human life and cultures is compatible
with a substantial decrease of human population. The flourishing of nonhuman life requires such a decrease."
Naess explained this [12] as: "Ecological egalitarianism implies the reinterpretation of the future-research
variable, "level of crowding," so that general mammalian crowding and loss of life-equality is taken seriously,
not only human crowding. (Research on the high requirements of free space of certain mammals has,
incidentally, suggested that theorists of human urbanism have largely underestimated human life-space
requirements. Behavioral crowding symptoms [neuroses, aggressiveness, loss of traditions] are largely the
same among mammals.)"

This has been seen by eco-feminists as "a grab at women’s special potency"

The tone of this article seems to me to be trying to provoke a response, particularly an aggressive one from
men. My simple response is that men are also part of nature with their own role to play, and if even a quick
reading of Taoism fails to see how it addresses male and female, then, the reader must be starting from a
very biased viewpoint.

There should not be a problem or difference between these various "eco" movements. That there is a conflict
seems indicative of the current stresses on society and individuals. There seems a need for "I’m right, your
wrong" type of statement, and for people who perhaps feel that they haven’t any power to be listened to and
to get a reaction that justifies themselves.

Alfred Alder, referred to by Metzner [4], believed that conscious feelings of superiority are always a
compensation for an unconscious inferiority complex and that such inferiority feelings tend to arise normally
in childhood, as a result of prolonged dependency and immaturity. Metzner [4], goes on to quote Paul Shepard
essay "a Post-historic primitivism" "incomplete, ontogeny runs into the dead end of immaturity and a miasma
of pathological limbos." (Ontogenesis = origin & development of an individual) as I interpret this it is
psychological speak for incomplete personal development leads to immaturity. The interesting part is what
causes this immaturity, and the resulting behaviour of the current dominant western culture. Shepard
proposes, I think if I understand his language, that it is the disconnection from the hunter-gatherer life and
hence from Nature, This disconnection is particularly critical at adolescence, and this is where the global
culture becomes a "laddish" culture, as noticed by Helena Norberg-Hodge in Learning from Ladakh.

4.0 Gaia Hypothesis

The Gaia Hypothesis takes the idea of systems further and applies it to the whole planet. All of life on earth
can be seen as whole that is more than the sum of its parts, this whole being like a huge super-lifeform that
we call 'Gaia' (after the name for the ancient Greek goddess of the earth). Living systems have a tendency
to keep themselves in balance but also to adapt and evolve over time. Scientists have found that the earth
also has these tendencies, with feedback mechanisms to 'keep in balance' the temperature and oxygen levels
of the atmosphere, just as our bodies maintain the temperature and oxygen levels in our arteries. The Gaia
Hypothesis is stating that the earth is alive and that we are part of it.

The Gaia hypothesis proposed by James Lovelock considers the Earth as a planet sized entity with properties
that could not be predicted from the sum of its parts. The reasons for this as stated in [6] are:

"Life first appeared on Earth about 3,500 million years ago. From that time until now, the presence of fossils
shows that the earth’s climate has changed very little. Yet the output of heat from the Sun, the surface
properties of the Earth, and the composition of the atmosphere have almost certainly varied greatly over the
same period.

The chemical composition of the atmosphere bears no relation to the expectations of steady- state chemical
equilibrium. The presence of methane, nitrous oxide, and even nitrogen in our present oxidising atmosphere
represents violation of the rules of chemistry to be measured in tens of orders of magnitude. Disequilibria on

55
this scale suggest that the atmosphere is not merely a biological product, but more probably a biological
construction: not living, but like a cat’s fur, a bird’s feathers, or the paper of a wasp’s nest, an extension of a
living system designed to maintain a chosen environment. Thus the atmospheric concentration of gases such
as oxygen and ammonia is found to be kept at an optimum value from which even small departures could
have disastrous consequences for life.

The climate and chemical properties of the Earth now and throughout its history seem always to have been
optimal for life…"

The use of an identity, especially a female one, appeals to a wider audience who can add anthropomorphic
attributes in much the same way a sailor might do to his or her ship. In this way it can help individuals to come
into a relationship with the planet and to foster the poetry and emotion that can enthuse people to care about
the issues that the "dry" scientific facts are presenting for our consideration.

To quote from James Lovelock[6] "It is an alternative to a pessimistic view that sees nature as a force to be
subdued and conquered. It is also an alternative to that equally depressing picture of our planet as a
demented spaceship, forever travelling, driverless and purposeless, around an inner circle of the sun."

James Lovelock says in [7] "Gaia theory provokes a view of the Earth where

Life is a planetary-scale phenomenon. On this scale it is near immortal and has no need to reproduce. There
can be no partial occupation of a planet by living organisms. It would be as impermanent as half an animal.
The presence of sufficient living organisms on a planet is needed for the regulation of the environment…

Theoretical ecology is enlarged. By tacking the species and their physical environment together as a single
system, we can …, build ecological models that are mathematically stable and yet include large numbers of
competing species..."

56
Technical Summary

TS.1 Framing and Context an overshoot. Overshoot pathways are characterized by the peak
magnitude of the overshoot, which may have implications for
impacts. All 1.5°C pathways involve limiting cumulative emissions
This chapter frames the context, knowledge-base and assessment of long-lived greenhouse gases, including carbon dioxide and nitrous
approaches used to understand the impacts of 1.5°C global warming oxide, and substantial reductions in other climate forcers (high
above pre-industrial levels and related global greenhouse gas confidence). Limiting cumulative emissions requires either reducing
emission pathways, building on the IPCC Fifth Assessment Report net global emissions of long-lived greenhouse gases to zero before
(AR5), in the context of strengthening the global response to the the cumulative limit is reached, or net negative global emissions
threat of climate change, sustainable development and efforts to (anthropogenic removals) after the limit is exceeded. {1.2.3, 1.2.4,
eradicate poverty. Cross-Chapter Boxes 1 and 2}

Human-induced warming reached approximately 1°C (likely) This report assesses projected impacts at a global average
between 0.8°C and 1.2°C) above pre-industrial levels in 2017, warming of 1.5°C and higher levels of warming. Global warming
increasing at 0.2°C (likely between 0.1°C and 0.3°C) per of 1.5°C is associated with global average surface temperatures
decade (high confidence). Global warming is defined in this report fluctuating naturally on either side of 1.5°C, together with warming
as an increase in combined surface air and sea surface temperatures substantially greater than 1.5°C in many regions and seasons (high TS
averaged over the globe and over a 30-year period. Unless otherwise confidence), all of which must be considered in the assessment of
specified, warming is expressed relative to the period 1850–1900, impacts. Impacts at 1.5°C of warming also depend on the emission
used as an approximation of pre-industrial temperatures in AR5. pathway to 1.5°C. Very different impacts result from pathways
For periods shorter than 30 years, warming refers to the estimated that remain below 1.5°C versus pathways that return to 1.5°C
average temperature over the 30 years centred on that shorter after a substantial overshoot, and when temperatures stabilize at
period, accounting for the impact of any temperature fluctuations 1.5°C versus a transient warming past 1.5°C (medium confidence).
or trend within those 30 years. Accordingly, warming from pre- {1.2.3, 1.3}
industrial levels to the decade 2006–2015 is assessed to be 0.87°C
(likely between 0.75°C and 0.99°C). Since 2000, the estimated level Ethical considerations, and the principle of equity in particular,
of human-induced warming has been equal to the level of observed are central to this report, recognizing that many of the impacts
warming with a likely range of ±20% accounting for uncertainty due of warming up to and beyond 1.5°C, and some potential
to contributions from solar and volcanic activity over the historical impacts of mitigation actions required to limit warming to
period (high confidence). {1.2.1} 1.5°C, fall disproportionately on the poor and vulnerable (high
confidence). Equity has procedural and distributive dimensions and
Warming greater than the global average has already been requires fairness in burden sharing both between generations and
experienced in many regions and seasons, with higher average between and within nations. In framing the objective of holding the
warming over land than over the ocean (high confidence). Most increase in the global average temperature rise to well below 2°C
land regions are experiencing greater warming than the global average, above pre-industrial levels, and to pursue efforts to limit warming to
while most ocean regions are warming at a slower rate. Depending 1.5°C, the Paris Agreement associates the principle of equity with the
on the temperature dataset considered, 20–40% of the global human broader goals of poverty eradication and sustainable development,
population live in regions that, by the decade 2006–2015, had already recognising that effective responses to climate change require a
experienced warming of more than 1.5°C above pre-industrial in at global collective effort that may be guided by the 2015 United
least one season (medium confidence). {1.2.1, 1.2.2} Nations Sustainable Development Goals. {1.1.1}

Past emissions alone are unlikely to raise global-mean Climate adaptation refers to the actions taken to manage
temperature to 1.5°C above pre-industrial levels (medium impacts of climate change by reducing vulnerability and
confidence), but past emissions do commit to other exposure to its harmful effects and exploiting any potential
changes, such as further sea level rise (high confidence). If all benefits. Adaptation takes place at international, national and
anthropogenic emissions (including aerosol-related) were reduced local levels. Subnational jurisdictions and entities, including urban
to zero immediately, any further warming beyond the 1°C already and rural municipalities, are key to developing and reinforcing
experienced would likely be less than 0.5°C over the next two to measures for reducing weather- and climate-related risks. Adaptation
three decades (high confidence), and likely less than 0.5°C on a implementation faces several barriers including lack of up-to-date and
century time scale (medium confidence), due to the opposing effects locally relevant information, lack of finance and technology, social
of different climate processes and drivers. A warming greater than values and attitudes, and institutional constraints (high confidence).
1.5°C is therefore not geophysically unavoidable: whether it will Adaptation is more likely to contribute to sustainable development
occur depends on future rates of emission reductions. {1.2.3, 1.2.4} when policies align with mitigation and poverty eradication goals
(medium confidence). {1.1, 1.4}
1.5°C emission pathways are defined as those that, given
current knowledge of the climate response, provide a one- Ambitious mitigation actions are indispensable to limit
in-two to two-in-three chance of warming either remaining warming to 1.5°C while achieving sustainable development
below 1.5°C or returning to 1.5°C by around 2100 following and poverty eradication (high confidence). Ill-designed responses,

57 31
Technical Summary

however, could pose challenges especially – but not exclusively – for TS.2 Mitigation Pathways Compatible
countries and regions contending with poverty and those requiring with 1.5°C in the Context of
significant transformation of their energy systems. This report focuses Sustainable Development
on ‘climate-resilient development pathways’, which aim to meet the
goals of sustainable development, including climate adaptation and
mitigation, poverty eradication and reducing inequalities. But any This chapter assesses mitigation pathways consistent with limiting
feasible pathway that remains within 1.5°C involves synergies and warming to 1.5°C above pre-industrial levels. In doing so, it explores
trade-offs (high confidence). Significant uncertainty remains as to the following key questions: What role do CO2 and non-CO2 emissions
which pathways are more consistent with the principle of equity. play? {2.2, 2.3, 2.4, 2.6} To what extent do 1.5°C pathways involve
{1.1.1, 1.4} overshooting and returning below 1.5°C during the 21st century? {2.2,
2.3} What are the implications for transitions in energy, land use and
Multiple forms of knowledge, including scientific evidence, sustainable development? {2.3, 2.4, 2.5} How do policy frameworks
narrative scenarios and prospective pathways, inform the affect the ability to limit warming to 1.5°C? {2.3, 2.5} What are the
understanding of 1.5°C. This report is informed by traditional associated knowledge gaps? {2.6}
evidence of the physical climate system and associated impacts and
TS vulnerabilities of climate change, together with knowledge drawn The assessed pathways describe integrated, quantitative
from the perceptions of risk and the experiences of climate impacts evolutions of all emissions over the 21st century associated
and governance systems. Scenarios and pathways are used to with global energy and land use and the world economy. The
explore conditions enabling goal-oriented futures while recognizing assessment is contingent upon available integrated assessment
the significance of ethical considerations, the principle of equity, and literature and model assumptions, and is complemented by other
the societal transformation needed. {1.2.3, 1.5.2} studies with different scope, for example, those focusing on individual
sectors. In recent years, integrated mitigation studies have improved
There is no single answer to the question of whether it the characterizations of mitigation pathways. However, limitations
is feasible to limit warming to 1.5°C and adapt to the remain, as climate damages, avoided impacts, or societal co-benefits
consequences. Feasibility is considered in this report as the of the modelled transformations remain largely unaccounted for, while
capacity of a system as a whole to achieve a specific outcome. The concurrent rapid technological changes, behavioural aspects, and
global transformation that would be needed to limit warming to uncertainties about input data present continuous challenges. (high
1.5°C requires enabling conditions that reflect the links, synergies confidence) {2.1.3, 2.3, 2.5.1, 2.6, Technical Annex 2}
and trade-offs between mitigation, adaptation and sustainable
development. These enabling conditions are assessed across many The Chances of Limiting Warming to 1.5°C
dimensions of feasibility – geophysical, environmental-ecological, and the Requirements for Urgent Action
technological, economic, socio-cultural and institutional – that
may be considered through the unifying lens of the Anthropocene, Pathways consistent with 1.5°C of warming above pre-industrial
acknowledging profound, differential but increasingly geologically levels can be identified under a range of assumptions about
significant human influences on the Earth system as a whole. This economic growth, technology developments and lifestyles.
framing also emphasises the global interconnectivity of past, present However, lack of global cooperation, lack of governance of the required
and future human–environment relations, highlighting the need and energy and land transformation, and increases in resource-intensive
opportunities for integrated responses to achieve the goals of the consumption are key impediments to achieving 1.5°C pathways.
Paris Agreement. {1.1, Cross-Chapter Box 1} Governance challenges have been related to scenarios with high
inequality and high population growth in the 1.5°C pathway literature.
{2.3.1, 2.3.2, 2.5}

Under emissions in line with current pledges under the Paris


Agreement (known as Nationally Determined Contributions,
or NDCs), global warming is expected to surpass 1.5°C above
pre-industrial levels, even if these pledges are supplemented
with very challenging increases in the scale and ambition of
mitigation after 2030 (high confidence). This increased action
would need to achieve net zero CO2 emissions in less than 15 years.
Even if this is achieved, temperatures would only be expected to remain
below the 1.5°C threshold if the actual geophysical response ends up
being towards the low end of the currently estimated uncertainty range.
Transition challenges as well as identified trade-offs can be reduced if
global emissions peak before 2030 and marked emissions reductions
compared to today are already achieved by 2030. {2.2, 2.3.5, Cross-
Chapter Box 11 in Chapter 4}

32 58
Technical Summary

Limiting warming to 1.5°C depends on greenhouse gas (GHG) 830 billion USD2010 (range of 150 billion to 1700 billion USD2010
emissions over the next decades, where lower GHG emissions in across six models). Total energy-related investments increase by about
2030 lead to a higher chance of keeping peak warming to 1.5°C 12% (range of 3% to 24%) in 1.5°C pathways relative to 2°C pathways.
(high confidence). Available pathways that aim for no or limited (less Average annual investment in low-carbon energy technologies and
than 0.1°C) overshoot of 1.5°C keep GHG emissions in 2030 to 25–30 energy efficiency are upscaled by roughly a factor of six (range of factor
GtCO2e yr−1 in 2030 (interquartile range). This contrasts with median of 4 to 10) by 2050 compared to 2015, overtaking fossil investments
estimates for current unconditional NDCs of 52–58 GtCO2e yr−1 in globally by around 2025 (medium confidence). Uncertainties and
2030. Pathways that aim for limiting warming to 1.5°C by 2100 after strategic mitigation portfolio choices affect the magnitude and focus
a temporary temperature overshoot rely on large-scale deployment of required investments. {2.5.2}
of carbon dioxide removal (CDR) measures, which are uncertain and
entail clear risks. In model pathways with no or limited overshoot of Future Emissions in 1.5°C Pathways
1.5°C, global net anthropogenic CO2 emissions decline by about 45%
from 2010 levels by 2030 (40–60% interquartile range), reaching net Mitigation requirements can be quantified using carbon budget
zero around 2050 (2045–2055 interquartile range). For limiting global approaches that relate cumulative CO2 emissions to global mean
warming to below 2°C with at least 66% probability CO2 emissions temperature increase. Robust physical understanding underpins
are projected to decline by about 25% by 2030 in most pathways (10– this relationship, but uncertainties become increasingly relevant as a TS
30% interquartile range) and reach net zero around 2070 (2065–2080 specific temperature limit is approached. These uncertainties relate to
interquartile range).1 {2.2, 2.3.3, 2.3.5, 2.5.3, Cross-Chapter Boxes 6 in the transient climate response to cumulative carbon emissions (TCRE),
Chapter 3 and 9 in Chapter 4, 4.3.7} non-CO2 emissions, radiative forcing and response, potential additional
Earth system feedbacks (such as permafrost thawing), and historical
Limiting warming to 1.5°C implies reaching net zero CO2 emissions and temperature. {2.2.2, 2.6.1}
emissions globally around 2050 and concurrent deep reductions
in emissions of non-CO2 forcers, particularly methane (high Cumulative CO2 emissions are kept within a budget by reducing
confidence). Such mitigation pathways are characterized by energy- global annual CO2 emissions to net zero. This assessment
demand reductions, decarbonization of electricity and other fuels, suggests a remaining budget of about 420 GtCO2 for a two-
electrification of energy end use, deep reductions in agricultural thirds chance of limiting warming to 1.5°C, and of about 580
emissions, and some form of CDR with carbon storage on land or GtCO2 for an even chance (medium confidence). The remaining
sequestration in geological reservoirs. Low energy demand and low carbon budget is defined here as cumulative CO2 emissions from the
demand for land- and GHG-intensive consumption goods facilitate start of 2018 until the time of net zero global emissions for global
limiting warming to as close as possible to 1.5°C. {2.2.2, 2.3.1, 2.3.5, warming defined as a change in global near-surface air temperatures.
2.5.1, Cross-Chapter Box 9 in Chapter 4}. Remaining budgets applicable to 2100 would be approximately
100 GtCO2 lower than this to account for permafrost thawing and
In comparison to a 2°C limit, the transformations required to limit potential methane release from wetlands in the future, and more
warming to 1.5°C are qualitatively similar but more pronounced thereafter. These estimates come with an additional geophysical
and rapid over the next decades (high confidence). 1.5°C implies uncertainty of at least ±400 GtCO2, related to non-CO2 response
very ambitious, internationally cooperative policy environments that and TCRE distribution. Uncertainties in the level of historic warming
transform both supply and demand (high confidence). {2.3, 2.4, 2.5} contribute ±250 GtCO2. In addition, these estimates can vary by
±250 GtCO2 depending on non-CO2 mitigation strategies as found in
Policies reflecting a high price on emissions are necessary available pathways. {2.2.2, 2.6.1}
in models to achieve cost-effective 1.5°C pathways (high
confidence). Other things being equal, modelling studies suggest Staying within a remaining carbon budget of 580 GtCO2 implies
the global average discounted marginal abatement costs for limiting that CO2 emissions reach carbon neutrality in about 30 years,
warming to 1.5°C being about 3–4 times higher compared to 2°C reduced to 20 years for a 420 GtCO2 remaining carbon budget
over the 21st century, with large variations across models and socio- (high confidence). The ±400 GtCO2 geophysical uncertainty range
economic and policy assumptions. Carbon pricing can be imposed surrounding a carbon budget translates into a variation of this timing
directly or implicitly by regulatory policies. Policy instruments, like of carbon neutrality of roughly ±15–20 years. If emissions do not start
technology policies or performance standards, can complement explicit declining in the next decade, the point of carbon neutrality would need
carbon pricing in specific areas. {2.5.1, 2.5.2, 4.4.5} to be reached at least two decades earlier to remain within the same
carbon budget. {2.2.2, 2.3.5}
Limiting warming to 1.5°C requires a marked shift in investment
patterns (medium confidence). Additional annual average energy- Non-CO2 emissions contribute to peak warming and thus
related investments for the period 2016 to 2050 in pathways limiting affect the remaining carbon budget. The evolution of
warming to 1.5°C compared to pathways without new climate policies methane and sulphur dioxide emissions strongly influences
beyond those in place today (i.e., baseline) are estimated to be around the chances of limiting warming to 1.5°C. In the near-term, a

1
Kyoto-GHG emissions in this statement are aggregated with GWP-100 values of the IPCC Second Assessment Report.

59 33
Technical Summary

weakening of aerosol cooling would add to future warming, equivalence method) supply a share of 52–67% (interquartile range)
but can be tempered by reductions in methane emissions (high of primary energy in 1.5°C pathways with no or limited overshoot;
confidence). Uncertainty in radiative forcing estimates (particularly while the share from coal decreases to 1–7% (interquartile range),
aerosol) affects carbon budgets and the certainty of pathway with a large fraction of this coal use combined with carbon capture
categorizations. Some non-CO2 forcers are emitted alongside CO2, and storage (CCS). From 2020 to 2050 the primary energy supplied
particularly in the energy and transport sectors, and can be largely by oil declines in most pathways (−39 to −77% interquartile range).
addressed through CO2 mitigation. Others require specific measures, Natural gas changes by −13% to −62% (interquartile range), but
for example, to target agricultural nitrous oxide (N2O) and methane some pathways show a marked increase albeit with widespread
(CH4), some sources of black carbon, or hydrofluorocarbons (high deployment of CCS. The overall deployment of CCS varies widely
confidence). In many cases, non-CO2 emissions reductions are similar across 1.5°C pathways with no or limited overshoot, with cumulative
in 2°C pathways, indicating reductions near their assumed maximum CO2 stored through 2050 ranging from zero up to 300 GtCO2
potential by integrated assessment models. Emissions of N2O and (minimum–maximum range), of which zero up to 140 GtCO2 is stored
NH3 increase in some pathways with strongly increased bioenergy from biomass. Primary energy supplied by bioenergy ranges from
demand. {2.2.2, 2.3.1, 2.4.2, 2.5.3} 40–310 EJ yr−1 in 2050 (minimum-maximum range), and nuclear from
3–66 EJ yr−1 (minimum–maximum range). These ranges reflect both
TS The Role of Carbon Dioxide Removal (CDR) uncertainties in technological development and strategic mitigation
portfolio choices. {2.4.2}
All analysed pathways limiting warming to 1.5°C with no
or limited overshoot use CDR to some extent to neutralize 1.5°C pathways with no or limited overshoot include a rapid
emissions from sources for which no mitigation measures decline in the carbon intensity of electricity and an increase
have been identified and, in most cases, also to achieve in electrification of energy end use (high confidence). By 2050,
net negative emissions to return global warming to 1.5°C the carbon intensity of electricity decreases to −92 to +11 gCO2 MJ−1
following a peak (high confidence). The longer the delay in (minimum–maximum range) from about 140 gCO2 MJ−1 in 2020,
reducing CO2 emissions towards zero, the larger the likelihood and electricity covers 34–71% (minimum–maximum range) of final
of exceeding 1.5°C, and the heavier the implied reliance on energy across 1.5°C pathways with no or limited overshoot from
net negative emissions after mid-century to return warming to about 20% in 2020. By 2050, the share of electricity supplied by
1.5°C (high confidence). The faster reduction of net CO2 emissions renewables increases to 59–97% (minimum-maximum range) across
in 1.5°C compared to 2°C pathways is predominantly achieved by 1.5°C pathways with no or limited overshoot. Pathways with higher
measures that result in less CO2 being produced and emitted, and chances of holding warming to below 1.5°C generally show a faster
only to a smaller degree through additional CDR. Limitations on decline in the carbon intensity of electricity by 2030 than pathways
the speed, scale and societal acceptability of CDR deployment also that temporarily overshoot 1.5°C. {2.4.1, 2.4.2, 2.4.3}
limit the conceivable extent of temperature overshoot. Limits to our
understanding of how the carbon cycle responds to net negative Transitions in global and regional land use are found in all
emissions increase the uncertainty about the effectiveness of CDR to pathways limiting global warming to 1.5°C with no or limited
decline temperatures after a peak. {2.2, 2.3, 2.6, 4.3.7} overshoot, but their scale depends on the pursued mitigation
portfolio (high confidence). Pathways that limit global warming to
CDR deployed at scale is unproven, and reliance on such 1.5°C with no or limited overshoot project a 4 million km2 reduction
technology is a major risk in the ability to limit warming to to a 2.5 million km2 increase of non-pasture agricultural land for food
1.5°C. CDR is needed less in pathways with particularly strong and feed crops and a 0.5–11 million km2 reduction of pasture land,
emphasis on energy efficiency and low demand. The scale and to be converted into 0-6 million km2 of agricultural land for energy
type of CDR deployment varies widely across 1.5°C pathways, crops and a 2 million km2 reduction to 9.5 million km2 increase in
with different consequences for achieving sustainable forests by 2050 relative to 2010 (medium confidence). Land-use
development objectives (high confidence). Some pathways rely transitions of similar magnitude can be observed in modelled 2°C
more on bioenergy with carbon capture and storage (BECCS), while pathways (medium confidence). Such large transitions pose profound
others rely more on afforestation, which are the two CDR methods challenges for sustainable management of the various demands on
most often included in integrated pathways. Trade-offs with other land for human settlements, food, livestock feed, fibre, bioenergy,
sustainability objectives occur predominantly through increased land, carbon storage, biodiversity and other ecosystem services (high
energy, water and investment demand. Bioenergy use is substantial confidence). {2.3.4, 2.4.4}
in 1.5°C pathways with or without BECCS due to its multiple roles in
decarbonizing energy use. {2.3.1, 2.5.3, 2.6.3, 4.3.7} Demand-Side Mitigation and Behavioural Changes

Properties of Energy and Land Transitions in 1.5°C Pathways Demand-side measures are key elements of 1.5°C pathways.
Lifestyle choices lowering energy demand and the land- and
The share of primary energy from renewables increases while GHG-intensity of food consumption can further support
coal usage decreases across pathways limiting warming to achievement of 1.5°C pathways (high confidence). By 2030 and
1.5°C with no or limited overshoot (high confidence). By 2050, 2050, all end-use sectors (including building, transport, and industry)
renewables (including bioenergy, hydro, wind, and solar, with direct- show marked energy demand reductions in modelled 1.5°C pathways,

34 60
Technical Summary

comparable and beyond those projected in 2°C pathways. Sectoral TS.3 Impacts of 1.5ºC Global Warming
models support the scale of these reductions. {2.3.4, 2.4.3, 2.5.1} on Natural and Human Systems
Links between 1.5°C Pathways and Sustainable Development
This chapter builds on findings of AR5 and assesses new scientific
Choices about mitigation portfolios for limiting warming to evidence of changes in the climate system and the associated impacts
1.5°C can positively or negatively impact the achievement of on natural and human systems, with a specific focus on the magnitude
other societal objectives, such as sustainable development and pattern of risks linked for global warming of 1.5°C above
(high confidence). In particular, demand-side and efficiency temperatures in the pre-industrial period. Chapter 3 explores observed
measures, and lifestyle choices that limit energy, resource, and impacts and projected risks to a range of natural and human systems,
GHG-intensive food demand support sustainable development with a focus on how risk levels change from 1.5°C to 2°C of global
(medium confidence). Limiting warming to 1.5°C can be achieved warming. The chapter also revisits major categories of risk (Reasons for
synergistically with poverty alleviation and improved energy security Concern, RFC) based on the assessment of new knowledge that has
and can provide large public health benefits through improved air become available since AR5.
quality, preventing millions of premature deaths. However, specific
mitigation measures, such as bioenergy, may result in trade-offs that 1.5°C and 2°C Warmer Worlds TS
require consideration. {2.5.1, 2.5.2, 2.5.3}
The global climate has changed relative to the pre-industrial
period, and there are multiple lines of evidence that these
changes have had impacts on organisms and ecosystems, as
well as on human systems and well-being (high confidence). The
increase in global mean surface temperature (GMST), which reached
0.87°C in 2006–2015 relative to 1850–1900, has increased the
frequency and magnitude of impacts (high confidence), strengthening
evidence of how an increase in GMST of 1.5°C or more could impact
natural and human systems (1.5°C versus 2°C). {3.3, 3.4, 3.5, 3.6,
Cross-Chapter Boxes 6, 7 and 8 in this chapter}

Human-induced global warming has already caused multiple


observed changes in the climate system (high confidence).
Changes include increases in both land and ocean temperatures, as well
as more frequent heatwaves in most land regions (high confidence).
There is also high confidence that global warming has resulted in an
increase in the frequency and duration of marine heatwaves. Further,
there is substantial evidence that human-induced global warming has
led to an increase in the frequency, intensity and/or amount of heavy
precipitation events at the global scale (medium confidence), as well
as an increased risk of drought in the Mediterranean region (medium
confidence). {3.3.1, 3.3.2, 3.3.3, 3.3.4, Box 3.4}

Trends in intensity and frequency of some climate and weather


extremes have been detected over time spans during which
about 0.5°C of global warming occurred (medium confidence).
This assessment is based on several lines of evidence, including
attribution studies for changes in extremes since 1950. {3.2, 3.3.1,
3.3.2, 3.3.3, 3.3.4}

Several regional changes in climate are assessed to occur with


global warming up to 1.5°C as compared to pre-industrial
levels, including warming of extreme temperatures in many
regions (high confidence), increases in frequency, intensity and/or
amount of heavy precipitation in several regions (high confidence),
and an increase in intensity or frequency of droughts in some regions
(medium confidence). {3.3.1, 3.3.2, 3.3.3, 3.3.4, Table 3.2}

There is no single ‘1.5°C warmer world’ (high confidence). In


addition to the overall increase in GMST, it is important to consider the

61 35
Technical Summary

size and duration of potential overshoots in temperature. Furthermore, and about 65 million fewer people being exposed to exceptional
there are questions on how the stabilization of an increase in GMST of heatwaves, assuming constant vulnerability (medium confidence).
1.5°C can be achieved, and how policies might be able to influence the {3.3.1, 3.3.2, Cross-Chapter Box 8 in this chapter}
resilience of human and natural systems, and the nature of regional
and subregional risks. Overshooting poses large risks for natural and Limiting global warming to 1.5°C would limit risks of increases
human systems, especially if the temperature at peak warming is in heavy precipitation events on a global scale and in several
high, because some risks may be long-lasting and irreversible, such regions compared to conditions at 2°C global warming
as the loss of some ecosystems (high confidence). The rate of change (medium confidence). The regions with the largest increases in heavy
for several types of risks may also have relevance, with potentially precipitation events for 1.5°C to 2°C global warming include: several
large risks in the case of a rapid rise to overshooting temperatures, high-latitude regions (e.g. Alaska/western Canada, eastern Canada/
even if a decrease to 1.5°C can be achieved at the end of the 21st Greenland/Iceland, northern Europe and northern Asia); mountainous
century or later (medium confidence). If overshoot is to be minimized, regions (e.g., Tibetan Plateau); eastern Asia (including China and Japan);
the remaining equivalent CO2 budget available for emissions is very and eastern North America (medium confidence). Tropical cyclones are
small, which implies that large, immediate and unprecedented global projected to decrease in frequency but with an increase in the number
efforts to mitigate greenhouse gases are required (high confidence). of very intense cyclones (limited evidence, low confidence). Heavy
TS {3.2, 3.6.2, Cross-Chapter Box 8 in this chapter} precipitation associated with tropical cyclones is projected to be higher
at 2°C compared to 1.5°C of global warming (medium confidence).
Robust1 global differences in temperature means and extremes Heavy precipitation, when aggregated at a global scale, is projected to
are expected if global warming reaches 1.5°C versus 2°C above be higher at 2°C than at 1.5°C of global warming (medium confidence)
the pre-industrial levels (high confidence). For oceans, regional {3.3.3, 3.3.6}
surface temperature means and extremes are projected to be higher
at 2°C compared to 1.5°C of global warming (high confidence). Limiting global warming to 1.5°C is expected to substantially
Temperature means and extremes are also projected to be higher at reduce the probability of extreme drought, precipitation deficits,
2°C compared to 1.5°C in most land regions, with increases being and risks associated with water availability (i.e., water stress) in
2–3 times greater than the increase in GMST projected for some some regions (medium confidence). In particular, risks associated
regions (high confidence). Robust increases in temperature means and with increases in drought frequency and magnitude are projected to be
extremes are also projected at 1.5°C compared to present-day values substantially larger at 2°C than at 1.5°C in the Mediterranean region
(high confidence) {3.3.1, 3.3.2}. There are decreases in the occurrence (including southern Europe, northern Africa and the Near East) and
of cold extremes, but substantial increases in their temperature, in southern Africa (medium confidence). {3.3.3, 3.3.4, Box 3.1, Box 3.2}
particular in regions with snow or ice cover (high confidence) {3.3.1}.
Risks to natural and human systems are expected to be lower
Climate models project robust2 differences in regional climate at 1.5°C than at 2°C of global warming (high confidence). This
between present-day and global warming up to 1.5°C3, and difference is due to the smaller rates and magnitudes of climate
between 1.5°C and 2°C3 (high confidence), depending on the change associated with a 1.5°C temperature increase, including lower
variable and region in question (high confidence). Large, robust frequencies and intensities of temperature-related extremes. Lower
and widespread differences are expected for temperature rates of change enhance the ability of natural and human systems
extremes (high confidence). Regarding hot extremes, the strongest to adapt, with substantial benefits for a wide range of terrestrial,
warming is expected to occur at mid-latitudes in the warm season (with freshwater, wetland, coastal and ocean ecosystems (including coral
increases of up to 3°C for 1.5°C of global warming, i.e., a factor of two) reefs) (high confidence), as well as food production systems, human
and at high latitudes in the cold season (with increases of up to 4.5°C health, and tourism (medium confidence), together with energy
at 1.5°C of global warming, i.e., a factor of three) (high confidence). systems and transportation (low confidence). {3.3.1, 3.4}
The strongest warming of hot extremes is projected to occur in
central and eastern North America, central and southern Europe, the Exposure to multiple and compound climate-related risks is
Mediterranean region (including southern Europe, northern Africa and projected to increase between 1.5°C and 2°C of global warming
the Near East), western and central Asia, and southern Africa (medium with greater proportions of people both exposed and susceptible to
confidence). The number of exceptionally hot days are expected to poverty in Africa and Asia (high confidence). For global warming from
increase the most in the tropics, where interannual temperature 1.5°C to 2°C, risks across energy, food, and water sectors could overlap
variability is lowest; extreme heatwaves are thus projected to emerge spatially and temporally, creating new – and exacerbating current –
earliest in these regions, and they are expected to already become hazards, exposures, and vulnerabilities that could affect increasing
widespread there at 1.5°C global warming (high confidence). Limiting numbers of people and regions (medium confidence). Small island
global warming to 1.5°C instead of 2°C could result in around 420 states and economically disadvantaged populations are particularly at
million fewer people being frequently exposed to extreme heatwaves, risk (high confidence). {3.3.1, 3.4.5.3, 3.4.5.6, 3.4.11, 3.5.4.9, Box 3.5}

2
Robust is used here to mean that at least two thirds of climate models show the same sign of changes at the grid point scale, and that differences in large regions are
statistically significant.
3
Projected changes in impacts between different levels of global warming are determined with respect to changes in global mean near-surface air temperature.

36 62
Technical Summary

Global warming of 2°C would lead to an expansion of areas with Future risks at 1.5°C of global warming will depend on the
significant increases in runoff, as well as those affected by flood mitigation pathway and on the possible occurrence of a
hazard, compared to conditions at 1.5°C (medium confidence). transient overshoot (high confidence). The impacts on natural
Global warming of 1.5°C would also lead to an expansion of the global and human systems would be greater if mitigation pathways
land area with significant increases in runoff (medium confidence) and temporarily overshoot 1.5°C and return to 1.5°C later in the century,
an increase in flood hazard in some regions (medium confidence) as compared to pathways that stabilize at 1.5°C without an overshoot
compared to present-day conditions. {3.3.5} (high confidence). The size and duration of an overshoot would also
affect future impacts (e.g., irreversible loss of some ecosystems) (high
The probability of a sea-ice-free Arctic Ocean4 during summer confidence). Changes in land use resulting from mitigation choices
is substantially higher at 2°C compared to 1.5°C of global could have impacts on food production and ecosystem diversity. {3.6.1,
warming (medium confidence). Model simulations suggest that 3.6.2, Cross-Chapter Boxes 7 and 8 in this chapter}
at least one sea-ice-free Arctic summer is expected every 10 years
for global warming of 2°C, with the frequency decreasing to one Climate Change Risks for Natural and Human systems
sea-ice-free Arctic summer every 100 years under 1.5°C (medium
confidence). An intermediate temperature overshoot will have no long- Terrestrial and Wetland Ecosystems
term consequences for Arctic sea ice coverage, and hysteresis is not TS
expected (high confidence). {3.3.8, 3.4.4.7} Risks of local species losses and, consequently, risks of
extinction are much less in a 1.5°C versus a 2°C warmer world
Global mean sea level rise (GMSLR) is projected to be around (high confidence). The number of species projected to lose over
0.1 m (0.04 – 0.16 m) less by the end of the 21st century in a half of their climatically determined geographic range at 2°C global
1.5°C warmer world compared to a 2°C warmer world (medium warming (18% of insects, 16% of plants, 8% of vertebrates) is
confidence). Projected GMSLR for 1.5°C of global warming has an projected to be reduced to 6% of insects, 8% of plants and 4% of
indicative range of 0.26 – 0.77m, relative to 1986–2005, (medium vertebrates at 1.5°C warming (medium confidence). Risks associated
confidence). A smaller sea level rise could mean that up to 10.4 million with other biodiversity-related factors, such as forest fires, extreme
fewer people (based on the 2010 global population and assuming no weather events, and the spread of invasive species, pests and
adaptation) would be exposed to the impacts of sea level rise globally diseases, would also be lower at 1.5°C than at 2°C of warming (high
in 2100 at 1.5°C compared to at 2°C. A slower rate of sea level rise confidence), supporting a greater persistence of ecosystem services.
enables greater opportunities for adaptation (medium confidence). {3.4.3, 3.5.2}
There is high confidence that sea level rise will continue beyond 2100.
Instabilities exist for both the Greenland and Antarctic ice sheets, which Constraining global warming to 1.5°C, rather than to 2°C
could result in multi-meter rises in sea level on time scales of century and higher, is projected to have many benefits for terrestrial
to millennia. There is medium confidence that these instabilities could and wetland ecosystems and for the preservation of their
be triggered at around 1.5°C to 2°C of global warming. {3.3.9, 3.4.5, services to humans (high confidence). Risks for natural and
3.6.3} managed ecosystems are higher on drylands compared to humid
lands. The global terrestrial land area projected to be affected by
The ocean has absorbed about 30% of the anthropogenic ecosystem transformations (13%, interquartile range 8–20%) at 2°C
carbon dioxide, resulting in ocean acidification and changes to is approximately halved at 1.5°C global warming to 4% (interquartile
carbonate chemistry that are unprecedented for at least the range 2–7%) (medium confidence). Above 1.5°C, an expansion of
last 65 million years (high confidence). Risks have been identified desert terrain and vegetation would occur in the Mediterranean
for the survival, calcification, growth, development and abundance of biome (medium confidence), causing changes unparalleled in the last
a broad range of marine taxonomic groups, ranging from algae to fish, 10,000 years (medium confidence). {3.3.2.2, 3.4.3.2, 3.4.3.5, 3.4.6.1,
with substantial evidence of predictable trait-based sensitivities (high 3.5.5.10, Box 4.2}
confidence). There are multiple lines of evidence that ocean warming
and acidification corresponding to 1.5°C of global warming would Many impacts are projected to be larger at higher latitudes,
impact a wide range of marine organisms and ecosystems, as well as owing to mean and cold-season warming rates above the
sectors such as aquaculture and fisheries (high confidence). {3.3.10, global average (medium confidence). High-latitude tundra and
3.4.4} boreal forest are particularly at risk, and woody shrubs are already
encroaching into tundra (high confidence) and will proceed with
Larger risks are expected for many regions and systems for further warming. Constraining warming to 1.5°C would prevent the
global warming at 1.5°C, as compared to today, with adaptation thawing of an estimated permafrost area of 1.5 to 2.5 million km2
required now and up to 1.5°C. However, risks would be larger at 2°C of over centuries compared to thawing under 2°C (medium confidence).
warming and an even greater effort would be needed for adaptation to {3.3.2, 3.4.3, 3.4.4}
a temperature increase of that magnitude (high confidence). {3.4, Box
3.4, Box 3.5, Cross-Chapter Box 6 in this chapter}

4
Ice free is defined for the Special Report as when the sea ice extent is less than 106 km2. Ice coverage less than this is considered to be equivalent to an ice-free Arctic Ocean
for practical purposes in all recent studies.

63 37
Technical Summary

Ocean Ecosystems (high confidence). A loss of 7–10% of rangeland livestock globally


is projected for approximately 2°C of warming, with considerable
Ocean ecosystems are already experiencing large-scale economic consequences for many communities and regions (medium
changes, and critical thresholds are expected to be reached at confidence). {3.4.6, 3.6, Box 3.1, Cross-Chapter Box 6 in this chapter}
1.5°C and higher levels of global warming (high confidence).
In the transition to 1.5°C of warming, changes to water temperatures Reductions in projected food availability are larger at 2°C
are expected to drive some species (e.g., plankton, fish) to relocate than at 1.5°C of global warming in the Sahel, southern Africa,
to higher latitudes and cause novel ecosystems to assemble (high the Mediterranean, central Europe and the Amazon (medium
confidence). Other ecosystems (e.g., kelp forests, coral reefs) are confidence). This suggests a transition from medium to high risk of
relatively less able to move, however, and are projected to experience regionally differentiated impacts on food security between 1.5°C and
high rates of mortality and loss (very high confidence). For example, 2°C (medium confidence). Future economic and trade environments
multiple lines of evidence indicate that the majority (70–90%) of and their response to changing food availability (medium confidence)
warm water (tropical) coral reefs that exist today will disappear even are important potential adaptation options for reducing hunger risk
if global warming is constrained to 1.5°C (very high confidence). in low- and middle-income countries. {Cross-Chapter Box 6 in this
{3.4.4, Box 3.4} chapter}
TS
Current ecosystem services from the ocean are expected to be Fisheries and aquaculture are important to global food security
reduced at 1.5°C of global warming, with losses being even but are already facing increasing risks from ocean warming
greater at 2°C of global warming (high confidence). The risks and acidification (medium confidence). These risks are
of declining ocean productivity, shifts of species to higher latitudes, projected to increase at 1.5°C of global warming and impact
damage to ecosystems (e.g., coral reefs, and mangroves, seagrass key organisms such as fin fish and bivalves (e.g., oysters),
and other wetland ecosystems), loss of fisheries productivity (at especially at low latitudes (medium confidence). Small-scale
low latitudes), and changes to ocean chemistry (e.g., acidification, fisheries in tropical regions, which are very dependent on habitat
hypoxia and dead zones) are projected to be substantially lower provided by coastal ecosystems such as coral reefs, mangroves,
when global warming is limited to 1.5°C (high confidence). {3.4.4, seagrass and kelp forests, are expected to face growing risks at 1.5°C
Box 3.4} of warming because of loss of habitat (medium confidence). Risks
of impacts and decreasing food security are projected to become
Water Resources greater as global warming reaches beyond 1.5°C and both ocean
warming and acidification increase, with substantial losses likely for
The projected frequency and magnitude of floods and droughts coastal livelihoods and industries (e.g., fisheries and aquaculture)
in some regions are smaller under 1.5°C than under 2°C of (medium to high confidence). {3.4.4, 3.4.5, 3.4.6, Box 3.1, Box 3.4,
warming (medium confidence). Human exposure to increased Box 3.5, Cross-Chapter Box 6 in this chapter}
flooding is projected to be substantially lower at 1.5°C compared to
2°C of global warming, although projected changes create regionally Land use and land-use change emerge as critical features of
differentiated risks (medium confidence). The differences in the risks virtually all mitigation pathways that seek to limit global
among regions are strongly influenced by local socio-economic warming to 1.5°C (high confidence). Most least-cost mitigation
conditions (medium confidence). {3.3.4, 3.3.5, 3.4.2} pathways to limit peak or end-of-century warming to 1.5°C make
use of carbon dioxide removal (CDR), predominantly employing
Risks of water scarcity are projected to be greater at 2°C than at significant levels of bioenergy with carbon capture and storage
1.5°C of global warming in some regions (medium confidence). (BECCS) and/or afforestation and reforestation (AR) in their portfolio
Depending on future socio-economic conditions, limiting global of mitigation measures (high confidence). {Cross-Chapter Box 7 in
warming to 1.5°C, compared to 2°C, may reduce the proportion of this chapter}
the world population exposed to a climate change-induced increase
in water stress by up to 50%, although there is considerable variability Large-scale deployment of BECCS and/or AR would have
between regions (medium confidence). Regions with particularly a far-reaching land and water footprint (high confidence).
large benefits could include the Mediterranean and the Caribbean Whether this footprint would result in adverse impacts, for example
(medium confidence). Socio-economic drivers, however, are expected on biodiversity or food production, depends on the existence and
to have a greater influence on these risks than the changes in climate effectiveness of measures to conserve land carbon stocks, measures
(medium confidence). {3.3.5, 3.4.2, Box 3.5} to limit agricultural expansion in order to protect natural ecosystems,
and the potential to increase agricultural productivity (medium
Land Use, Food Security and Food Production Systems agreement). In addition, BECCS and/or AR would have substantial
direct effects on regional climate through biophysical feedbacks,
Limiting global warming to 1.5°C, compared with 2°C, is which are generally not included in Integrated Assessments Models
projected to result in smaller net reductions in yields of maize, (high confidence). {3.6.2, Cross-Chapter Boxes 7 and 8 in this chapter}
rice, wheat, and potentially other cereal crops, particularly in
sub-Saharan Africa, Southeast Asia, and Central and South America; The impacts of large-scale CDR deployment could be greatly
and in the CO2-dependent nutritional quality of rice and wheat reduced if a wider portfolio of CDR options were deployed, if a

38 64
Technical Summary

holistic policy for sustainable land management were adopted, change should global warming increase from 1.5°C to 2°C (medium
and if increased mitigation efforts were employed to strongly confidence). {3.5}
limit the demand for land, energy and material resources,
including through lifestyle and dietary changes (medium Global warming has already affected tourism, with increased
confidence). In particular, reforestation could be associated with risks projected under 1.5°C of warming in specific geographic
significant co-benefits if implemented in a manner than helps restore regions and for seasonal tourism including sun, beach and
natural ecosystems (high confidence). {Cross-Chapter Box 7 in this snow sports destinations (very high confidence). Risks will be
chapter} lower for tourism markets that are less climate sensitive, such as
gaming and large hotel-based activities (high confidence). Risks for
Human Health, Well-Being, Cities and Poverty coastal tourism, particularly in subtropical and tropical regions, will
increase with temperature-related degradation (e.g., heat extremes,
Any increase in global temperature (e.g., +0.5°C) is projected storms) or loss of beach and coral reef assets (high confidence).
to affect human health, with primarily negative consequences {3.3.6, 3.4.4.12, 3.4.9.1, Box 3.4}
(high confidence). Lower risks are projected at 1.5°C than at 2°C
for heat-related morbidity and mortality (very high confidence), and Small Islands, and Coastal and Low-lying areas
for ozone-related mortality if emissions needed for ozone formation TS
remain high (high confidence). Urban heat islands often amplify the Small islands are projected to experience multiple inter-
impacts of heatwaves in cities (high confidence). Risks for some related risks at 1.5°C of global warming that will increase with
vector-borne diseases, such as malaria and dengue fever are projected warming of 2°C and higher levels (high confidence). Climate
to increase with warming from 1.5°C to 2°C, including potential hazards at 1.5°C are projected to be lower compared to those at 2°C
shifts in their geographic range (high confidence). Overall for vector- (high confidence). Long-term risks of coastal flooding and impacts on
borne diseases, whether projections are positive or negative depends populations, infrastructures and assets (high confidence), freshwater
on the disease, region and extent of change (high confidence). Lower stress (medium confidence), and risks across marine ecosystems (high
risks of undernutrition are projected at 1.5°C than at 2°C (medium confidence) and critical sectors (medium confidence) are projected to
confidence). Incorporating estimates of adaptation into projections increase at 1.5°C compared to present-day levels and increase further
reduces the magnitude of risks (high confidence). {3.4.7, 3.4.7.1, at 2°C, limiting adaptation opportunities and increasing loss and
3.4.8, 3.5.5.8} damage (medium confidence). Migration in small islands (internally
and internationally) occurs for multiple reasons and purposes, mostly
Global warming of 2°C is expected to pose greater risks to urban for better livelihood opportunities (high confidence) and increasingly
areas than global warming of 1.5°C (medium confidence). The owing to sea level rise (medium confidence). {3.3.2.2, 3.3.6–9,
extent of risk depends on human vulnerability and the effectiveness 3.4.3.2, 3.4.4.2, 3.4.4.5, 3.4.4.12, 3.4.5.3, 3.4.7.1, 3.4.9.1, 3.5.4.9,
of adaptation for regions (coastal and non-coastal), informal Box 3.4, Box 3.5}
settlements and infrastructure sectors (such as energy, water and
transport) (high confidence). {3.4.5, 3.4.8} Impacts associated with sea level rise and changes to the
salinity of coastal groundwater, increased flooding and damage
Poverty and disadvantage have increased with recent warming to infrastructure, are projected to be critically important in
(about 1°C) and are expected to increase for many populations vulnerable environments, such as small islands, low-lying
as average global temperatures increase from 1°C to 1.5°C coasts and deltas, at global warming of 1.5°C and 2°C (high
and higher (medium confidence). Outmigration in agricultural- confidence). Localized subsidence and changes to river discharge can
dependent communities is positively and statistically significantly potentially exacerbate these effects. Adaptation is already happening
associated with global temperature (medium confidence). Our (high confidence) and will remain important over multi-centennial
understanding of the links of 1.5°C and 2°C of global warming to time scales. {3.4.5.3, 3.4.5.4, 3.4.5.7, 5.4.5.4, Box 3.5}
human migration are limited and represent an important knowledge
gap. {3.4.10, 3.4.11, 5.2.2, Table 3.5} Existing and restored natural coastal ecosystems may be
effective in reducing the adverse impacts of rising sea levels
Key Economic Sectors and Services and intensifying storms by protecting coastal and deltaic
regions (medium confidence). Natural sedimentation rates are
Risks to global aggregated economic growth due to climate expected to be able to offset the effect of rising sea levels, given
change impacts are projected to be lower at 1.5°C than at 2°C the slower rates of sea level rise associated with 1.5°C of warming
by the end of this century (medium confidence). {3.5.2, 3.5.3} (medium confidence). Other feedbacks, such as landward migration
of wetlands and the adaptation of infrastructure, remain important
The largest reductions in economic growth at 2°C compared (medium confidence). {3.4.4.12, 3.4.5.4, 3.4.5.7}
to 1.5°C of warming are projected for low- and middle-income
countries and regions (the African continent, Southeast Asia, Increased Reasons for Concern
India, Brazil and Mexico) (low to medium confidence). Countries
in the tropics and Southern Hemisphere subtropics are projected to There are multiple lines of evidence that since AR5 the assessed
experience the largest impacts on economic growth due to climate levels of risk increased for four of the five Reasons for Concern

65 39
Technical Summary

(RFCs) for global warming levels of up to 2°C (high confidence). TS.4 Strengthening and Implementing
The risk transitions by degrees of global warming are now: from high the Global Response
to very high between 1.5°C and 2°C for RFC1 (Unique and threatened
systems) (high confidence); from moderate to high risk between 1°C and
1.5°C for RFC2 (Extreme weather events) (medium confidence); from Limiting warming to 1.5°C above pre-industrial levels would
moderate to high risk between 1.5°C and 2°C for RFC3 (Distribution of require transformative systemic change, integrated with
impacts) (high confidence); from moderate to high risk between 1.5°C sustainable development. Such change would require the
and 2.5°C for RFC4 (Global aggregate impacts) (medium confidence); upscaling and acceleration of the implementation of far-
and from moderate to high risk between 1°C and 2.5°C for RFC5 reaching, multilevel and cross-sectoral climate mitigation
(Large-scale singular events) (medium confidence). {3.5.2} and addressing barriers. Such systemic change would need
to be linked to complementary adaptation actions, including
1. The category ‘Unique and threatened systems’ (RFC1) transformational adaptation, especially for pathways that
display a transition from high to very high risk which is temporarily overshoot 1.5°C (medium evidence, high agreement)
now located between 1.5°C and 2°C of global warming as {Chapter 2, Chapter 3, 4.2.1, 4.4.5, 4.5}. Current national pledges
opposed to at 2.6°C of global warming in AR5, owing to new and on mitigation and adaptation are not enough to stay below the Paris
TS multiple lines of evidence for changing risks for coral reefs, the Agreement temperature limits and achieve its adaptation goals. While
Arctic and biodiversity in general (high confidence). {3.5.2.1} transitions in energy efficiency, carbon intensity of fuels, electrification
and land-use change are underway in various countries, limiting
2. In ‘Extreme weather events’ (RFC2), the transition from warming to 1.5°C will require a greater scale and pace of change to
moderate to high risk is now located between 1.0°C and transform energy, land, urban and industrial systems globally. {4.3, 4.4,
1.5°C of global warming, which is very similar to the AR5 Cross-Chapter Box 9 in this Chapter}
assessment but is projected with greater confidence (medium
confidence). The impact literature contains little information Although multiple communities around the world are
about the potential for human society to adapt to extreme demonstrating the possibility of implementation consistent with
weather events, and hence it has not been possible to locate 1.5°C pathways {Boxes 4.1-4.10}, very few countries, regions,
the transition from ‘high’ to ‘very high’ risk within the context of cities, communities or businesses can currently make such
assessing impacts at 1.5°C versus 2°C of global warming. There a claim (high confidence). To strengthen the global response,
is thus low confidence in the level at which global warming could almost all countries would need to significantly raise their level
lead to very high risks associated with extreme weather events in of ambition. Implementation of this raised ambition would
the context of this report. {3.5} require enhanced institutional capabilities in all countries,
including building the capability to utilize indigenous and
3. With respect to the ‘Distribution of impacts’ (RFC3) a local knowledge (medium evidence, high agreement). In developing
transition from moderate to high risk is now located countries and for poor and vulnerable people, implementing the
between 1.5°C and 2°C of global warming, compared with response would require financial, technological and other forms of
between 1.6°C and 2.6°C global warming in AR5, owing to new support to build capacity, for which additional local, national and
evidence about regionally differentiated risks to food security, international resources would need to be mobilized (high confidence).
water resources, drought, heat exposure and coastal submergence However, public, financial, institutional and innovation capabilities
(high confidence). {3.5} currently fall short of implementing far-reaching measures at scale in
all countries (high confidence). Transnational networks that support
4. In ‘global aggregate impacts’ (RFC4) a transition from multilevel climate action are growing, but challenges in their scale-up
moderate to high levels of risk is now located between remain. {4.4.1, 4.4.2, 4.4.4, 4.4.5, Box 4.1, Box 4.2, Box 4.7}
1.5°C and 2.5°C of global warming, as opposed to at 3.6°C of
warming in AR5, owing to new evidence about global aggregate Adaptation needs will be lower in a 1.5°C world compared to
economic impacts and risks to Earth’s biodiversity (medium a 2°C world (high confidence) {Chapter 3; Cross-Chapter Box 11
confidence). {3.5} in this chapter}. Learning from current adaptation practices and
strengthening them through adaptive governance {4.4.1}, lifestyle
5. Finally, ‘large-scale singular events’ (RFC5), moderate risk and behavioural change {4.4.3} and innovative financing mechanisms
is now located at 1°C of global warming and high risk is {4.4.5} can help their mainstreaming within sustainable development
located at 2.5°C of global warming, as opposed to at 1.6°C practices. Preventing maladaptation, drawing on bottom-up approaches
(moderate risk) and around 4°C (high risk) in AR5, because of new {Box 4.6} and using indigenous knowledge {Box 4.3} would effectively
observations and models of the West Antarctic ice sheet (medium engage and protect vulnerable people and communities. While
confidence). {3.3.9, 3.5.2, 3.6.3} adaptation finance has increased quantitatively, significant further
expansion would be needed to adapt to 1.5°C. Qualitative gaps in the
distribution of adaptation finance, readiness to absorb resources, and
monitoring mechanisms undermine the potential of adaptation finance
to reduce impacts. {Chapter 3, 4.4.2, 4.4.5, 4.6}

40 66
Technical Summary

System Transitions existing agricultural systems generally reduces the emissions intensity
of food production and offers strong synergies with rural development,
The energy system transition that would be required to limit poverty reduction and food security objectives, but options to reduce
global warming to 1.5°C above pre-industrial conditions is absolute emissions are limited unless paired with demand-side
underway in many sectors and regions around the world measures. Technological innovation including biotechnology, with
(medium evidence, high agreement). The political, economic, social adequate safeguards, could contribute to resolving current feasibility
and technical feasibility of solar energy, wind energy and electricity constraints and expand the future mitigation potential of agriculture.
storage technologies has improved dramatically over the past few {4.3.2, 4.4.4}
years, while that of nuclear energy and carbon dioxide capture
and storage (CCS) in the electricity sector have not shown similar Shifts in dietary choices towards foods with lower emissions
improvements. {4.3.1} and requirements for land, along with reduced food loss and
waste, could reduce emissions and increase adaptation options
Electrification, hydrogen, bio-based feedstocks and substitution, (high confidence). Decreasing food loss and waste and changing
and, in several cases, carbon dioxide capture, utilization and dietary behaviour could result in mitigation and adaptation (high
storage (CCUS) would lead to the deep emissions reductions confidence) by reducing both emissions and pressure on land, with
required in energy-intensive industries to limit warming to significant co-benefits for food security, human health and sustainable TS
1.5°C. However, those options are limited by institutional, economic and development {4.3.2, 4.4.5, 4.5.2, 4.5.3, 5.4.2}, but evidence of
technical constraints, which increase financial risks to many incumbent successful policies to modify dietary choices remains limited.
firms (medium evidence, high agreement). Energy efficiency in industry
is more economically feasible and helps enable industrial system Mitigation and Adaptation Options and Other Measures
transitions but would have to be complemented with greenhouse gas
(GHG)-neutral processes or carbon dioxide removal (CDR) to make A mix of mitigation and adaptation options implemented in a
energy-intensive industries consistent with 1.5°C (high confidence). participatory and integrated manner can enable rapid, systemic
{4.3.1, 4.3.4} transitions – in urban and rural areas – that are necessary
elements of an accelerated transition consistent with limiting
Global and regional land-use and ecosystems transitions and warming to 1.5°C. Such options and changes are most effective
associated changes in behaviour that would be required to when aligned with economic and sustainable development,
limit warming to 1.5°C can enhance future adaptation and and when local and regional governments are supported by
land-based agricultural and forestry mitigation potential. Such national governments {4.3.3, 4.4.1, 4.4.3}. Various mitigation
transitions could, however, carry consequences for livelihoods options are expanding rapidly across many geographies. Although
that depend on agriculture and natural resources {4.3.2, Cross- many have development synergies, not all income groups have so
Chapter Box 6 in Chapter 3}. Alterations of agriculture and forest far benefited from them. Electrification, end-use energy efficiency
systems to achieve mitigation goals could affect current ecosystems and increased share of renewables, amongst other options, are
and their services and potentially threaten food, water and livelihood lowering energy use and decarbonizing energy supply in the built
security. While this could limit the social and environmental feasibility environment, especially in buildings. Other rapid changes needed in
of land-based mitigation options, careful design and implementation urban environments include demotorization and decarbonization of
could enhance their acceptability and support sustainable development transport, including the expansion of electric vehicles, and greater use
objectives (medium evidence, medium agreement). {4.3.2, 4.5.3} of energy-efficient appliances (medium evidence, high agreement).
Technological and social innovations can contribute to limiting
Changing agricultural practices can be an effective climate warming to 1.5°C, for example, by enabling the use of smart grids,
adaptation strategy. A diversity of adaptation options exists, energy storage technologies and general-purpose technologies, such
including mixed crop-livestock production systems which can be a as information and communication technology (ICT) that can be
cost-effective adaptation strategy in many global agriculture systems deployed to help reduce emissions. Feasible adaptation options include
(robust evidence, medium agreement). Improving irrigation efficiency green infrastructure, resilient water and urban ecosystem services,
could effectively deal with changing global water endowments, urban and peri-urban agriculture, and adapting buildings and land use
especially if achieved via farmers adopting new behaviours and water- through regulation and planning (medium evidence, medium to high
efficient practices rather than through large-scale infrastructural agreement). {4.3.3, 4.4.3, 4.4.4}
interventions (medium evidence, medium agreement). Well-designed
adaptation processes such as community-based adaptation can be Synergies can be achieved across systemic transitions through
effective depending upon context and levels of vulnerability. {4.3.2, several overarching adaptation options in rural and urban areas.
4.5.3} Investments in health, social security and risk sharing and spreading
are cost-effective adaptation measures with high potential for scaling
Improving the efficiency of food production and closing yield up (medium evidence, medium to high agreement). Disaster risk
gaps have the potential to reduce emissions from agriculture, management and education-based adaptation have lower prospects of
reduce pressure on land, and enhance food security and future scalability and cost-effectiveness (medium evidence, high agreement)
mitigation potential (high confidence). Improving productivity of but are critical for building adaptive capacity. {4.3.5, 4.5.3}

67 41
Technical Summary

Converging adaptation and mitigation options can lead to understanding about their effectiveness to limit global warming; and
synergies and potentially increase cost-effectiveness, but a weak capacity to govern, legitimize, and scale such measures. Some
multiple trade-offs can limit the speed of and potential for recent model-based analysis suggests SRM would be effective but that
scaling up. Many examples of synergies and trade-offs exist in it is too early to evaluate its feasibility. Even in the uncertain case that
all sectors and system transitions. For instance, sustainable water the most adverse side-effects of SRM can be avoided, public resistance,
management (high evidence, medium agreement) and investment in ethical concerns and potential impacts on sustainable development
green infrastructure (medium evidence, high agreement) to deliver could render SRM economically, socially and institutionally undesirable
sustainable water and environmental services and to support urban (low agreement, medium evidence). {4.3.8, Cross-Chapter Box 10 in
agriculture are less cost-effective than other adaptation options but this chapter}
can help build climate resilience. Achieving the governance, finance
and social support required to enable these synergies and to avoid Enabling Rapid and Far-Reaching Change
trade-offs is often challenging, especially when addressing multiple
objectives, and attempting appropriate sequencing and timing of The speed of transitions and of technological change required
interventions. {4.3.2, 4.3.4, 4.4.1, 4.5.2, 4.5.3, 4.5.4} to limit warming to 1.5°C above pre-industrial levels has been
observed in the past within specific sectors and technologies
TS Though CO2 dominates long-term warming, the reduction of {4.2.2.1}. But the geographical and economic scales at which
warming short-lived climate forcers (SLCFs), such as methane the required rates of change in the energy, land, urban,
and black carbon, can in the short term contribute significantly to infrastructure and industrial systems would need to take place
limiting warming to 1.5°C above pre-industrial levels. Reductions are larger and have no documented historic precedent (limited
of black carbon and methane would have substantial co-benefits evidence, medium agreement). To reduce inequality and alleviate
(high confidence), including improved health due to reduced air poverty, such transformations would require more planning and
pollution. This, in turn, enhances the institutional and socio- stronger institutions (including inclusive markets) than observed in the
cultural feasibility of such actions. Reductions of several warming past, as well as stronger coordination and disruptive innovation across
SLCFs are constrained by economic and social feasibility (low evidence, actors and scales of governance. {4.3, 4.4}
high agreement). As they are often co-emitted with CO2, achieving the
energy, land and urban transitions necessary to limit warming to 1.5°C Governance consistent with limiting warming to 1.5°C and the
would see emissions of warming SLCFs greatly reduced. {2.3.3.2, 4.3.6} political economy of adaptation and mitigation can enable and
accelerate systems transitions,behavioural change,innovation and
Most CDR options face multiple feasibility constraints, which technology deployment (medium evidence, medium agreement).
differ between options, limiting the potential for any single For 1.5°C-consistent actions, an effective governance framework
option to sustainably achieve the large-scale deployment would include: accountable multilevel governance that includes non-
required in the 1.5°C-consistent pathways described in state actors, such as industry, civil society and scientific institutions;
Chapter 2 (high confidence). Those 1.5°C pathways typically rely coordinated sectoral and cross-sectoral policies that enable collaborative
on bioenergy with carbon capture and storage (BECCS), afforestation multi-stakeholder partnerships; strengthened global-to-local financial
and reforestation (AR), or both, to neutralize emissions that are architecture that enables greater access to finance and technology;
expensive to avoid, or to draw down CO2 emissions in excess of the addressing climate-related trade barriers; improved climate education
carbon budget {Chapter 2}. Though BECCS and AR may be technically and greater public awareness; arrangements to enable accelerated
and geophysically feasible, they face partially overlapping yet different behaviour change; strengthened climate monitoring and evaluation
constraints related to land use. The land footprint per tonne of CO2 systems; and reciprocal international agreements that are sensitive
removed is higher for AR than for BECCS, but given the low levels of to equity and the Sustainable Development Goals (SDGs). System
current deployment, the speed and scales required for limiting warming transitions can be enabled by enhancing the capacities of public, private
to 1.5°C pose a considerable implementation challenge, even if the and financial institutions to accelerate climate change policy planning
issues of public acceptance and absence of economic incentives were and implementation, along with accelerated technological innovation,
to be resolved (high agreement, medium evidence). The large potential deployment and upkeep. {4.4.1, 4.4.2, 4.4.3, 4.4.4}
of afforestation and the co-benefits if implemented appropriately (e.g.,
on biodiversity and soil quality) will diminish over time, as forests Behaviour change and demand-side management can
saturate (high confidence). The energy requirements and economic significantly reduce emissions, substantially limiting the
costs of direct air carbon capture and storage (DACCS) and enhanced reliance on CDR to limit warming to 1.5°C {Chapter 2, 4.4.3}.
weathering remain high (medium evidence, medium agreement). At the Political and financial stakeholders may find climate actions more cost-
local scale, soil carbon sequestration has co-benefits with agriculture effective and socially acceptable if multiple factors affecting behaviour
and is cost-effective even without climate policy (high confidence). Its are considered, including aligning these actions with people’s core
potential feasibility and cost-effectiveness at the global scale appears values (medium evidence, high agreement). Behaviour- and lifestyle-
to be more limited. {4.3.7} related measures and demand-side management have already led
to emission reductions around the world and can enable significant
Uncertainties surrounding solar radiation modification future reductions (high confidence). Social innovation through bottom-
(SRM) measures constrain their potential deployment. These up initiatives can result in greater participation in the governance of
uncertainties include: technological immaturity; limited physical systems transitions and increase support for technologies, practices

42 68
Technical Summary

and policies that are part of the global response to limit warming to would help redirect capital away from potentially stranded assets
1.5°C . {Chapter 2, 4.4.1, 4.4.3, Figure 4.3} (medium evidence, medium agreement). {4.4.5}

This rapid and far-reaching response required to keep warming Knowledge Gaps
below 1.5°C and enhance the capacity to adapt to climate risks
would require large increases of investments in low-emission Knowledge gaps around implementing and strengthening the
infrastructure and buildings, along with a redirection of financial global response to climate change would need to be urgently
flows towards low-emission investments (robust evidence, high resolved if the transition to a 1.5°C world is to become reality.
agreement). An estimated mean annual incremental investment of Remaining questions include: how much can be realistically expected
around 1.5% of global gross fixed capital formation (GFCF) for the from innovation and behavioural and systemic political and economic
energy sector is indicated between 2016 and 2035, as well as about changes in improving resilience, enhancing adaptation and reducing
2.5% of global GFCF for other development infrastructure that could GHG emissions? How can rates of changes be accelerated and scaled
also address SDG implementation. Though quality policy design and up? What is the outcome of realistic assessments of mitigation and
effective implementation may enhance efficiency, they cannot fully adaptation land transitions that are compliant with sustainable
substitute for these investments. {2.5.2, 4.2.1, 4.4.5} development, poverty eradication and addressing inequality? What are
life-cycle emissions and prospects of early-stage CDR options? How TS
Enabling this investment requires the mobilization and better can climate and sustainable development policies converge, and how
integration of a range of policy instruments that include the can they be organised within a global governance framework and
reduction of socially inefficient fossil fuel subsidy regimes and innovative financial system, based on principles of justice and ethics (including
price and non-price national and international policy instruments. These ‘common but differentiated responsibilities and respective capabilities’
would need to be complemented by de-risking financial instruments (CBDR-RC)), reciprocity and partnership? To what extent would
and the emergence of long-term low-emission assets. These instruments limiting warming to 1.5°C require a harmonization of macro-financial
would aim to reduce the demand for carbon-intensive services and shift and fiscal policies, which could include financial regulators such as
market preferences away from fossil fuel-based technology. Evidence central banks? How can different actors and processes in climate
and theory suggest that carbon pricing alone, in the absence of governance reinforce each other, and hedge against the fragmentation
sufficient transfers to compensate their unintended distributional cross- of initiatives? {4.1, 4.3.7, 4.4.1, 4.4.5, 4.6}
sector, cross-nation effects, cannot reach the incentive levels needed
to trigger system transitions (robust evidence, medium agreement).
But, embedded in consistent policy packages, they can help mobilize
incremental resources and provide flexible mechanisms that help reduce
the social and economic costs of the triggering phase of the transition
(robust evidence, medium agreement). {4.4.3, 4.4.4, 4.4.5}

Increasing evidence suggests that a climate-sensitive


realignment of savings and expenditure towards low-emission,
climate-resilient infrastructure and services requires an
evolution of global and national financial systems. Estimates
suggest that, in addition to climate-friendly allocation of public
investments, a potential redirection of 5% to 10% of the annual
capital revenues5 is necessary for limiting warming to 1.5°C {4.4.5,
Table 1 in Box 4.8}. This could be facilitated by a change of incentives
for private day-to-day expenditure and the redirection of savings
from speculative and precautionary investments towards long-
term productive low-emission assets and services. This implies the
mobilization of institutional investors and mainstreaming of climate
finance within financial and banking system regulation. Access by
developing countries to low-risk and low-interest finance through
multilateral and national development banks would have to be
facilitated (medium evidence, high agreement). New forms of public–
private partnerships may be needed with multilateral, sovereign and
sub-sovereign guarantees to de-risk climate-friendly investments,
support new business models for small-scale enterprises and help
households with limited access to capital. Ultimately, the aim is to
promote a portfolio shift towards long-term low-emission assets that

5
Annual capital revenues are the paid interests plus the increase of the asset value.

69 43
Technical Summary

TS.5 Sustainable Development, Poverty confidence). Many strategies for sustainable development enable
Eradication and Reducing Inequalities transformational adaptation for a 1.5°C warmer world, provided
attention is paid to reducing poverty in all its forms and to promoting
equity and participation in decision-making (medium evidence, high
This chapter takes sustainable development as the starting point and agreement). As such, sustainable development has the potential
focus for analysis. It considers the broad and multifaceted bi-directional to significantly reduce systemic vulnerability, enhance adaptive
interplay between sustainable development, including its focus on capacity, and promote livelihood security for poor and disadvantaged
eradicating poverty and reducing inequality in their multidimensional populations (high confidence). {5.3.1}
aspects, and climate actions in a 1.5°C warmer world. These fundamental
connections are embedded in the Sustainable Development Goals Synergies between adaptation strategies and the SDGs are
(SDGs). The chapter also examines synergies and trade-offs of expected to hold true in a 1.5°C warmer world, across sectors
adaptation and mitigation options with sustainable development and and contexts (medium evidence, medium agreement). Synergies
the SDGs and offers insights into possible pathways, especially climate- between adaptation and sustainable development are significant
resilient development pathways towards a 1.5°C warmer world. for agriculture and health, advancing SDGs 1 (extreme poverty),
2 (hunger), 3 (healthy lives and well-being) and 6 (clean water) (robust
TS Sustainable Development, Poverty and Inequality evidence, medium agreement). {5.3.2} Ecosystem- and community-
in a 1.5°C Warmer World based adaptation, along with the incorporation of indigenous and
local knowledge, advances synergies with SDGs 5 (gender equality),
Limiting global warming to 1.5°C rather than 2°C above pre- 10 (reducing inequalities) and 16 (inclusive societies), as exemplified
industrial levels would make it markedly easier to achieve many in drylands and the Arctic (high evidence, medium agreement). {5.3.2,
aspects of sustainable development, with greater potential to Box 5.1, Cross-Chapter Box 10 in Chapter 4}
eradicate poverty and reduce inequalities (medium evidence, high
agreement). Impacts avoided with the lower temperature limit could Adaptation strategies can result in trade-offs with and
reduce the number of people exposed to climate risks and vulnerable among the SDGs (medium evidence, high agreement). Strategies
to poverty by 62 to 457 million, and lessen the risks of poor people that advance one SDG may create negative consequences for other
to experience food and water insecurity, adverse health impacts, and SDGs, for instance SDGs 3 (health) versus 7 (energy consumption)
economic losses, particularly in regions that already face development and agricultural adaptation and SDG 2 (food security) versus SDGs 3
challenges (medium evidence, medium agreement). {5.2.2, 5.2.3} (health), 5 (gender equality), 6 (clean water), 10 (reducing inequalities),
Avoided impacts expected to occur between 1.5°C and 2°C warming 14 (life below water) and 15 (life on the land) (medium evidence,
would also make it easier to achieve certain SDGs, such as those that medium agreement). {5.3.2}
relate to poverty, hunger, health, water and sanitation, cities and
ecosystems (SDGs 1, 2, 3, 6, 11, 14 and 15) (medium evidence, high Pursuing place-specific adaptation pathways towards a
agreement). {5.2.3, Table 5.2 available at the end of the chapter} 1.5°C warmer world has the potential for significant positive
outcomes for well-being in countries at all levels of development
Compared to current conditions, 1.5°C of global warming would (medium evidence, high agreement). Positive outcomes emerge when
nonetheless pose heightened risks to eradicating poverty, adaptation pathways (i) ensure a diversity of adaptation options based
reducing inequalities and ensuring human and ecosystem well- on people’s values and the trade-offs they consider acceptable, (ii)
being (medium evidence, high agreement). Warming of 1.5°C is not maximize synergies with sustainable development through inclusive,
considered ‘safe’ for most nations, communities, ecosystems and participatory and deliberative processes, and (iii) facilitate equitable
sectors and poses significant risks to natural and human systems as transformation. Yet such pathways would be difficult to achieve
compared to the current warming of 1°C (high confidence). {Cross- without redistributive measures to overcome path dependencies,
Chapter Box 12 in Chapter 5} The impacts of 1.5°C of warming would uneven power structures, and entrenched social inequalities (medium
disproportionately affect disadvantaged and vulnerable populations evidence, high agreement). {5.3.3}
through food insecurity, higher food prices, income losses, lost
livelihood opportunities, adverse health impacts and population Mitigation and Sustainable Development
displacements (medium evidence, high agreement). {5.2.1} Some of
the worst impacts on sustainable development are expected to be The deployment of mitigation options consistent with 1.5°C
felt among agricultural and coastal dependent livelihoods, indigenous pathways leads to multiple synergies across a range of
people, children and the elderly, poor labourers, poor urban dwellers in sustainable development dimensions. At the same time, the
African cities, and people and ecosystems in the Arctic and Small Island rapid pace and magnitude of change that would be required to
Developing States (SIDS) (medium evidence, high agreement). {5.2.1, limit warming to 1.5°C, if not carefully managed, would lead to
Box 5.3, Chapter 3, Box 3.5, Cross-Chapter Box 9 in Chapter 4} trade-offs with some sustainable development dimensions (high
confidence). The number of synergies between mitigation response
Climate Adaptation and Sustainable Development options and sustainable development exceeds the number of trade-
offs in energy demand and supply sectors; agriculture, forestry and
Prioritization of sustainable development and meeting the other land use (AFOLU); and for oceans (very high confidence). {Figure
SDGs is consistent with efforts to adapt to climate change (high 5.2, Table 5.2 available at the end of the chapter} The 1.5°C pathways

44 70
Technical Summary

indicate robust synergies, particularly for the SDGs 3 (health), 7 (energy), promote diversification of the economy and the energy sector could
12 (responsible consumption and production) and 14 (oceans) (very ease this transition (medium evidence, high agreement). {5.4.1.2,
high confidence). {5.4.2, Figure 5.3} For SDGs 1 (poverty), 2 (hunger), Box 5.2}
6 (water) and 7 (energy), there is a risk of trade-offs or negative side
effects from stringent mitigation actions compatible with 1.5°C of Sustainable Development Pathways to 1.5°C
warming (medium evidence, high agreement). {5.4.2}
Sustainable development broadly supports and often enables
Appropriately designed mitigation actions to reduce energy the fundamental societal and systems transformations that
demand can advance multiple SDGs simultaneously. Pathways would be required for limiting warming to 1.5°C above pre-
compatible with 1.5°C that feature low energy demand show the industrial levels (high confidence). Simulated pathways that
most pronounced synergies and the lowest number of trade-offs feature the most sustainable worlds (e.g., Shared Socio-Economic
with respect to sustainable development and the SDGs (very high Pathways (SSP) 1) are associated with relatively lower mitigation and
confidence). Accelerating energy efficiency in all sectors has synergies adaptation challenges and limit warming to 1.5°C at comparatively
with SDGs 7 (energy), 9 (industry, innovation and infrastructure), lower mitigation costs. In contrast, development pathways with high
11 (sustainable cities and communities), 12 (responsible consumption fragmentation, inequality and poverty (e.g., SSP3) are associated with
and production), 16 (peace, justice and strong institutions), and comparatively higher mitigation and adaptation challenges. In such TS
17 (partnerships for the goals) (robust evidence, high agreement). pathways, it is not possible to limit warming to 1.5°C for the vast
{5.4.1, Figure 5.2, Table 5.2} Low-demand pathways, which would majority of the integrated assessment models (medium evidence,
reduce or completely avoid the reliance on bioenergy with carbon high agreement). {5.5.2} In all SSPs, mitigation costs substantially
capture and storage (BECCS) in 1.5°C pathways, would result in increase in 1.5°C pathways compared to 2°C pathways. No pathway
significantly reduced pressure on food security, lower food prices and in the literature integrates or achieves all 17 SDGs (high confidence).
fewer people at risk of hunger (medium evidence, high agreement). {5.5.2} Real-world experiences at the project level show that the
{5.4.2, Figure 5.3} actual integration between adaptation, mitigation and sustainable
development is challenging as it requires reconciling trade-offs across
The impacts of carbon dioxide removal options on SDGs depend sectors and spatial scales (very high confidence). {5.5.1}
on the type of options and the scale of deployment (high
confidence). If poorly implemented, carbon dioxide removal (CDR) Without societal transformation and rapid implementation
options such as bioenergy, BECCS and AFOLU would lead to trade- of ambitious greenhouse gas reduction measures, pathways
offs. Appropriate design and implementation requires considering to limiting warming to 1.5°C and achieving sustainable
local people’s needs, biodiversity and other sustainable development development will be exceedingly difficult, if not impossible,
dimensions (very high confidence). {5.4.1.3, Cross-Chapter Box 7 in to achieve (high confidence). The potential for pursuing such
Chapter 3} pathways differs between and within nations and regions, due to
different development trajectories, opportunities and challenges (very
The design of the mitigation portfolios and policy instruments high confidence). {5.5.3.2, Figure 5.1} Limiting warming to 1.5°C
to limit warming to 1.5°C will largely determine the overall would require all countries and non-state actors to strengthen their
synergies and trade-offs between mitigation and sustainable contributions without delay. This could be achieved through sharing
development (very high confidence). Redistributive policies that efforts based on bolder and more committed cooperation, with support
shield the poor and vulnerable can resolve trade-offs for a range for those with the least capacity to adapt, mitigate and transform
of SDGs (medium evidence, high agreement). Individual mitigation (medium evidence, high agreement). {5.5.3.1, 5.5.3.2} Current
options are associated with both positive and negative interactions efforts towards reconciling low-carbon trajectories and reducing
with the SDGs (very high confidence). {5.4.1} However, appropriate inequalities, including those that avoid difficult trade-offs associated
choices across the mitigation portfolio can help to maximize positive with transformation, are partially successful yet demonstrate notable
side effects while minimizing negative side effects (high confidence). obstacles (medium evidence, medium agreement). {5.5.3.3, Box 5.3,
{5.4.2, 5.5.2} Investment needs for complementary policies resolving Cross-Chapter Box 13 in this chapter}
trade-offs with a range of SDGs are only a small fraction of the overall
mitigation investments in 1.5°C pathways (medium evidence, high Social justice and equity are core aspects of climate-resilient
agreement). {5.4.2, Figure 5.4} Integration of mitigation with adaptation development pathways for transformational social change.
and sustainable development compatible with 1.5°C warming requires Addressing challenges and widening opportunities between
a systems perspective (high confidence). {5.4.2, 5.5.2} and within countries and communities would be necessary
to achieve sustainable development and limit warming to
Mitigation consistent with 1.5°C of warming create high risks 1.5°C, without making the poor and disadvantaged worse off
for sustainable development in countries with high dependency (high confidence). Identifying and navigating inclusive and socially
on fossil fuels for revenue and employment generation (high acceptable pathways towards low-carbon, climate-resilient futures is a
confidence). These risks are caused by the reduction of global demand challenging yet important endeavour, fraught with moral, practical and
affecting mining activity and export revenues and challenges to rapidly political difficulties and inevitable trade-offs (very high confidence).
decrease high carbon intensity of the domestic economy (robust {5.5.2, 5.5.3.3, Box 5.3} It entails deliberation and problem-solving
evidence, high agreement). {5.4.1.2, Box 5.2} Targeted policies that processes to negotiate societal values, well-being, risks and resilience

71 45
Technical Summary

and to determine what is desirable and fair, and to whom (medium


evidence, high agreement). Pathways that encompass joint, iterative
planning and transformative visions, for instance in Pacific SIDS
like Vanuatu and in urban contexts, show potential for liveable and
sustainable futures (high confidence). {5.5.3.1, 5.5.3.3, Figure 5.5,
Box 5.3, Cross-Chapter Box 13 in this chapter}

The fundamental societal and systemic changes to achieve


sustainable development, eradicate poverty and reduce
inequalities while limiting warming to 1.5°C would require
meeting a set of institutional, social, cultural, economic and
technological conditions (high confidence). The coordination
and monitoring of policy actions across sectors and spatial scales
is essential to support sustainable development in 1.5°C warmer
conditions (very high confidence). {5.6.2, Box 5.3} External funding
TS and technology transfer better support these efforts when they
consider recipients’ context-specific needs (medium evidence, high
agreement). {5.6.1} Inclusive processes can facilitate transformations
by ensuring participation, transparency, capacity building and iterative
social learning (high confidence). {5.5.3.3, Cross-Chapter Box 13,
5.6.3} Attention to power asymmetries and unequal opportunities
for development, among and within countries, is key to adopting
1.5°C-compatible development pathways that benefit all populations
(high confidence). {5.5.3, 5.6.4, Box 5.3} Re-examining individual and
collective values could help spur urgent, ambitious and cooperative
change (medium evidence, high agreement). {5.5.3, 5.6.5}

46 72
05 Understanding Climate Change Mitigation, Adaptation and Governance

1. Climate Change Mitigation:


Mitigation (of climate change) is a human intervention to reduce emissions or enhance the sinks of
greenhouse gases.

Mitigation Options
Energy system transition: A long-term structural change from a high-carbon energy system to a low-
carbon energy system. For example, switch to wind (on-shore & off-shore), solar, bioenergy, electricity
storage, power sector carbon dioxide capture and storage, and nuclear energy.

Settlement & infrastructure system transition: Changes in land-use & urban planning, shift to electric
cars and buses, sharing schemes, public transport, non-motorized transport, aviation & shipping, smart
grids, efficient appliances, and low/zero-energy buildings.

Land & ecosystem transition: Reduced food wastage & efficient food production, dietary shifts,
sustainable intensification of agriculture, and ecosystems restoration.
Afforestation: Planting of new forests on lands that historically have not contained forests.

Reforestation: Planting of forests on lands that have previously contained forests but that have been
converted to some other use.

Reducing Emissions from Deforestation and Forest Degradation (REDD+): An effort to create
financial value for the carbon stored in forests, offering incentives for developing countries to reduce
emissions from forested lands and invest in low-carbon paths to sustainable development (SD). It is
therefore a mechanism for mitigation that results from avoiding deforestation. REDD+ goes beyond
deforestation and forest degradation, and includes the role of conservation, sustainable management of
forests and enhancement of forest carbon stocks.

Bioenergy: Energy derived from any form of biomass or its metabolic by-products. Some pathways rely
more on bioenergy with carbon capture and storage (BECCS), while others rely more on afforestation,
which are the two CDR methods most often included in integrated pathways.

Carbon Geoengineering/ Carbon dioxide removal (CDR): Anthropogenic activities removing CO2
from the atmosphere and durably storing it in geological, terrestrial, or ocean reservoirs, or in products. It
includes existing and potential anthropogenic enhancement of biological or geochemical sinks and direct
air capture and storage, but excludes natural CO2 uptake not directly caused by human activities. This
involves Carbon sequestration, Carbon dioxide capture and storage (CCS) and Carbon dioxide capture
and utilisation (CCU). Most least-cost mitigation pathways to limit peak or end-of-century warming to
1.5°C make use of carbon dioxide removal (CDR), predominantly employing significant levels of
bioenergy with carbon capture and storage (BECCS) and/or afforestation and reforestation (AR) in their
portfolio of mitigation measures (high confidence).

Carbon dioxide capture and storage (CCS): A process in which a relatively pure stream of carbon
dioxide (CO2) from industrial and energy-related sources is separated (captured), conditioned,
compressed and transported to a storage location for long-term isolation from the atmosphere.
Sometimes referred to as Carbon capture and storage.

Carbon dioxide capture and utilisation (CCU): A process in which CO2 is captured and then used to
produce a new product. If the CO2 is stored in a product for a climate-relevant time horizon, this is
referred to as carbon dioxide capture, utilisation and storage (CCUS). Only then, and only combined with
CO2 recently removed from the atmosphere, can CCUS lead to carbon dioxide removal. CCU is
sometimes referred to as carbon dioxide capture and use.

Energy efficiency: The goal of a given country, or73


the global community as a whole, to maintain an
adequate, stable and predictable energy supply. Measures encompass safeguarding the sufficiency of
energy resources to meet national energy demand at competitive and stable prices and the resilience of
the energy supply; enabling development and deployment of technologies; building sufficient
infrastructure to generate, store and transmit energy supplies; and ensuring enforceable contracts of
delivery.

Behaviour change and demand-side management: Behaviour change and demand-side


management can significantly reduce emissions, substantially limiting the reliance on CDR to limit
warming to 1.5°C.

Industrial system transition: Energy efficiency, bio-based & circularity, electrification & hydrogen, and
industrial carbon dioxide capture, utilization, and storage.

Decarbonisation: The process by which countries, individuals or other entities aim to achieve zero
fossil carbon existence. Typically refers to a reduction of the carbon emissions associated with
electricity, industry, and transport.

Table 1: Feasibility of 1.5 ºC- relevant mitigation options


System Mitigation Option Evidence Agreement Context
Wind regime economic status space for wind farms and the
Wind energy Robust Medium existence of a legal framework for independent power
(on-shore & off-shore) producers affect uptake; cost-effectiveness affected by incentive
regime
Cost effectiveness affected by solar irradiation and incentive
Solar PV Robust High regime. Also enhanced by legal framework for independent
power producers, which affects uptake
Energy Depends on availability of biomass and land and the capability
System Bioenergy Robust Medium to manage sustainable land use. Distributional effects depend
Transitions on the agrarian (or other) system used to produce feedstock.
Batteries universal, but grid flexible resources vary with area’s
Electricity Storage Robust High level of development.
Power sector carbon Varies with local CO2 storage capacity, presence of legal
dioxide capture and Robust High framework, level of development and quality of public
storage engagement
Electricity market organization legal framework, standardization
Nuclear energy Robust High & know how, country’s democratic fabric, institutional and
ethnical capacity and safety culture of public and private
institutions.
Reduced food wastage Will depend on the combination of individual and institutional
& efficient food Robust High behavior
Land & production
Ecosysem Depends on individual behavior education, cultural factors and
Transitions Dietary shifts Medium High institutional support

Sustainable Depends on development and deployment of new technologies.


intensification of Medium Hig
agriculture
Ecosystems restoration Depends on location and institutional factors
Medium High
Land use & urban Varies with urban fabric, not geography or economy; requires
planning Robust Medium capacitated local government and legitimate tenure system
Varies with degree of government intervention; requires
Electric Cars & Buses Medium High capacity to retrofit “fuelling” stations

Historic schemes universal but new ones depend on ICT status;


Urban & Sharing schemes Limited Medium undermined by high crime and low levels of law enforcement.
Infrastructure
System Depends on presence of existing ‘informal’ taxi systems which
Transitions Public Transport Robust High may be more cost-effective and affordable than capital intensive
new build schemes as well as(local) government capabilities
Varies with technology, governance and accountability
Aviation & shipping Medium Medium
Varies with economic status and presence or quality of existing
Smart grids Medium Medium grid
Adoption varies with economic status and policy framework
Efficient appliances Medium High
Low/zero-energy Depends on size of existing building stock and growth of
buildings Medium High building stock
Potential and adoption depend on existing efficiency, energy
Energy Efficiency Robust High prices and interest rates, as well as government incentives.

Faces barriers in terms of pressure on natural resources and


Bio-based & circularity Medium 74
Medium biodiversity product substitutions depends on market
organization and government incentivization.
System Mitigation Option Evidence Agreement Context
Industrial Depends on availability of large-scale cheap, emission-free
System Electrification & Medium High electricity (electrification, hydrogen) or CO2 storage nearby
Transitions hydrogen (hydrogen). Manufacturers’ appetite to embrace disruptive
innovations
Industrial carbon High Concentration of CO2, in exhaust gas improves economic
dioxide capture, Robust High and technical feasibility of CCUS in industry. CO2 storage or
utilization and storage reuse possibilities.
Bioenergy and carbon Depens on biomass availability, CO2 storage capacity, legal
dioxide capture and Medium Medium framework, economic status and social acceptance.
storage
Direct air carbon Depends on CO2 free energy, CO2 storage capacity legal
dioxide capture and Medium Medium framework, economic status and social acceptance.
Carbon storage
Dioxide Afforestation Depends on location, mode of implementation, and economic
Removal &reforestation Robust High and institutional factors
Soil carbon
Sequestration & bio Robust High Depends on location, soil properties, time span
char
Enhanced weathering Depends on CO2 free energy, economic status and social
Medium Low acceptance

2. Climate Change Adaptation


Adaptation: In human systems, the process of adjustment to actual or expected climate and its effects,
in order to moderate harm or exploit beneficial opportunities. In natural systems, the process of
adjustment to actual climate and its effects; human intervention may facilitate adjustment to expected
climate and its effects. Climate adaptation refers to the actions taken to manage impacts of climate
change by reducing vulnerability and exposure to its harmful effects and exploiting any potential
benefits. Adaptation implementation faces several barriers including lack of up-to-date and locally
relevant information, lack of finance and technology, social values and attitudes, and institutional
constraints (high confidence).

Adaptation pathways: A series of adaptation choices involving trade-offs between short-term and long-
term goals and values. These are processes of deliberation to identify solutions that are meaningful to
people in the context of their daily lives and to avoid potential maladaptation.

Options for transformational adaptations:


Transformational adaptation: Adaptation that changes the fundamental attributes of a socio-ecological
system in anticipation of climate change and its impacts.

Climate-smart agriculture (CSA): Climate-smart agriculture (CSA) is an approach that helps to guide
actions needed to transform and reorient agricultural systems to effectively support development and
ensure food security in a changing climate. CSA aims to tackle three main objectives: sustainably
increasing agricultural productivity and incomes, adapting and building resilience to climate change, and
reducing and/or removing greenhouse gas emissions, where possible (FAO 2018).

Conservation agriculture (CA): It is a set of soil management practices that minimize the disruption of
the soil's structure, composition and natural biodiversity. The three principles of CA are: minimum tillage
and soil disruption, permanent soil cover, and crop rotation. These practices increase soil organic
carbon. 1% increase in soil organic carbon equals sequestering 18 tons of CO2 in soil per acre.

Mixed crop-livestock production systems: Changing agricultural practices can be an effective climate
adaptation strategy. A diversity of adaptation options exists, including mixed crop-livestock production
systems, which can be a cost-effective adaptation strategy in many global agriculture systems (robust
evidence, medium agreement).

Ecosystem-based adaptations: Ecosystem-based adaptations are nature-based solutions that


harness biodiversity and ecosystem services to reduce vulnerability and build resilience to climate
change. Ecosystem-based adaptations involve the conservation, sustainable management, and
restoration of ecosystems can help people adapt to75
the impacts of climate change.
Community-based adaptations: Community-based adaptations to climate change are approaches that
aims to include vulnerable people in the design and implementation of climate change adaptation
measures.

Climate services: Climate services refers to information and products that enhance users’ knowledge
and understanding about the impacts of climate change and/or climate variability so as to aid decision-
making of individuals and organizations and enable preparedness and early climate change action.
Products can include climate data products.

Soil carbon sequestration (SCS): Land management changes, which increase the soil organic carbon
content, resulting in a net removal of CO2 from the atmosphere.

Soil moisture: Water stored in the soil in liquid or frozen form. Root- zone soil moisture is of most
relevance for plant activity. Higher soil organic carbon content helps in better soils moisture holding
capacity.

Urban & infrastructure system transition: Sustainable land-use & urban planning, sustainable water
management, green infrastructure & ecosystem services, and building codes & standards.

Adaptation finance: The finance that funds efforts to adapt to the impacts of climate change qualifies
as adaptation finance.

Adaptive governance: An emerging term in the literature for the evolution of formal and informal
institutions of governance that prioritize social learning in planning, implementation and evaluation of
policy through iterative social learning to steer the use and protection of natural resources, ecosystem
services and common pool natural resources, particularly in situations of complexity and uncertainty.

Indigenous knowledge: Indigenous knowledge operates at a much finer spatial and temporal scale
than science, and includes understandings of how to cope with and adapt to environmental variability
and trends.

Efficient livestock system: Livestock systems that are highly diverse and intensive production
systems, which can integrate livestock-crop-forage systems to improve feed efficiency and reduce time
from birth to harvest.

Socio-cultural acceptability: Acceptance of appropriate technology, methods, and systems by


overcoming social, cultural, and sometimes religious barriers.

Ecosystem-Based Approaches to Adaptation1

Ecosystem-based adaptation (EBA), defined as the use of biodiversity and ecosystem services as
part of an overall adaptation strategy to help people to adapt to the adverse effects of climate change
(CBD, 2009), integrates the use of biodiversity and ecosystem services into climate change
adaptation strategies (e.g., CBD, 2009; Munroe et al., 2011; EBA is implemented through the
sustainable management of natural resources and conservation and restoration of ecosystems, to
provide and sustain services that facilitate adaptation both to climate variability and change (Colls et
al., 2009). It also sets out to take into account the multiple social, economic, and cultural co-benefits
for local communities.

EBA can be combined with, or even serve as a substitute for, the use of engineered infrastructure or
other technological approaches. Engineered defenses such as dams, sea walls, and levees adversely
affect biodiversity, potentially resulting in maladaptation due to damage to ecosystem regulating
services (Campbell et al., 2009; Munroe et al., 2011).

1 Reproduced from IPCC (2014) Climate Change 2014: Impacts, Adaptation and Vulnerability
76
There is some evidence that the restoration and use of ecosystem services may reduce or delay the
need for these engineering solutions (CBD, 2009). EBA offers lower risk of maladaptation than
engineering solutions in that their application is more flexible and responsive to unanticipated
environmental changes.

Well-integrated EBA can be more cost effective and sustainable than non-integrated physical
engineering approaches (Jones et al., 2012), and may contribute to achieving sustainable
development goals (e.g., poverty reduction, sustainable environmental management, and even
mitigation objectives), especially when they are integrated with sound ecosystem management
approaches (CBD, 2009). In addition, EBA yields economic, social, and environmental co-benefits in
the form of ecosystem goods and services (World Bank, 2009).
• EBA is applicable in both developed and developing countries. In developing countries where
economies depend more directly on the provision of ecosystem services (Vignola et al.,
2009),
• EBA may be a highly useful approach to reduce risks to climate change impacts and ensure
that development proceeds on a pathways that are resilient to climate change (Munang et al.,
2013).
• EBA projects may be developed by enhancing existing initiatives, such as community-based
adaptation and natural resource management approaches (e.g., Khan et al., 2012, Midgley et
al., 2012; Roberts et al., 2012).

Examples of ecosystem based approaches to adaptation include:

• Sustainable water management, where river basins, aquifers, flood plains, and their
associated vegetation are managed or restored to provide resilient water storage and
enhanced baseflows, flood regulation and protection services, reduction of erosion/siltation
rates, and more ecosystem goods (e.g., Opperman et al., 2009; Midgley et al., 2012)
• Disaster risk reduction through the restoration of coastal habitats (e.g., mangroves, wetlands,
and deltas) to provide effective measure against storm-surges, saline intrusion, and coastal
erosion (Jonkman et al., 2013)
• Sustainable management of grasslands and rangelands to enhance pastoral livelihoods and
increase resilience to drought and flooding
• Establishment of diverse and resilient agricultural systems, and adapting crop and livestock
variety mixes to secure food provision. Traditional knowledge may contribute in this area
through, for example, identifying indigenous crop and livestock genetic diversity, and water
conservation techniques.
• Management of fire-prone ecosystems to achieve safer fire regimes while ensuring the
maintenance of natural processes

Table 2: Feasibility of 1.5 ºC- relevant adaptation options 77


System Adaptation Option Evidence Agreement Context

Power infrastructure, Depends on existing power infrastructure, all


including water Medium High generation sources and those with intensive water
requirements
Conservation Depends on irrigated/rain fed system, ecosystem
Agriculture Medium Medium characteristics, crop type, other farming practices
Efficient irrigation Depends on agricultural system technology used,
Limited High regional institutional and biophysical context
Efficient Livestock Dependent on livestock breeds, feed practices, and
Systems Limited High biophysical context (e.g. carrying capacity)
Depends on knowledge, financial support and market
Energy Agroforestry Medium High conditions
System Community based Focus on rural areas and combined with ecosystems-
Transitions adaptation Medium High based adaptation, does not include urban settings
Ecosystem restoration & Mostly focused on existing and evaluated REDD+
avoided deforestation Robust Medium projects.
Biodiversity Focus on hotspots of biodiversity vulnerability and
management Medium Medium high connectivity
Coastal defense & Depends on locations that require it as a first
hardening Robust Medium adaptation option
Sustainable aquaculture Depends on locations at risk and socio-cultural
Limited Medium context
Sustainable land-use & Depends on nature of planning systems and
urban planning Medium Medium enforcement mechanism
Sustainable Water Balancing sustainable water supply and rising
Urban & Management Robust Medium demand especially in low-income counties
Infrastructure Green infrastructure & Depends on reconciliation of urban development with
System ecosystem services Medium High green infrastructure
Transitions Building codes & Adoption requires legal, education and enforcement
standards Limited Medium mechanisms to regulate buildings
Industrial Intensive industry Depends on intensive industry, existing infrastructure
Ststen infrastructure resilience Limited High and using or requiring high demand of water.
Transitions and water management
Disaster risk Requires institutional, technical, and financial capacity
Management Medium High in frontline agencies and government.
Risk spreading and Requires well-developed financial structures and
sharing Medium Medium public understanding
Depends on climate information availability and
Overarching Climate Services Medium High usability, local infrastructure and institutions, national
Adaptation priorities
Options Indigenous knowledge Dependent on recognition of indigenous rights, laws,
Medium High and governance systems

Education & Learning Existing education system, funding


Medium High

Population health and Requires basis health services and infrastrures


health system Medium High

Type and mechanism of safety net, political priorities,


Social safety nets Medium Medium institutional transparency

Hazard exposure, political and socio-cultural


Human migration Medium Low acceptability (I destination), migrant skills and social
networks.

78
06 Key Green Business Concepts
Life Cycle Analysis
Life cycle assessment (LCA) Compilation and evaluation of the inputs, outputs and the potential
environmental impacts of a product or service throughout its life cycle. This definition builds from ISO
(2018).
LCA is a widely used technique defined by ISO 14040 as a “compilation and evaluation of the inputs,
outputs and the potential environmental impacts of a product system throughout its life cycle”. The results
of LCA studies are strongly dependent on the system boundaries within which they are conducted. The
technique is intended for relative comparison of two similar means to complete a product.
Life cycle analysis (or assessment) or LCA, is a method for evaluating the total environmental impact of a
product or service. The major work involved in conducting an LCA is to compile all the inputs and outputs
of a product or a system of products, and to evaluate the potential environmental impacts of a product or
system throughout its life cycle, from extraction of raw materials through manufacturing, use, and
disposal.
LCA is a methodological tool that helps calculate the total environmental costs (including externalities
such as pollution) of a particular product or service over its lifetime or to compare the costs of different
options by applying a life cycle perspective of activities related to processes or products. The life cycle
perspective focuses on the entire life cycle of a good or service, starting from the extraction and
processing of raw materials, followed by manufacturing, transportation and use, and ending with waste
management as well as recycling and final disposal. Each of these life cycle stages consumes energy
and resources, and generates emissions and wastes, which result in a number of environmental impacts.
Without LCA, many of these impacts may be overlooked.
A full LCA is sometimes called “cradle-to-grave” because it covers all phases from manufacture to
disposal. This type of LCA can provide information to help consumers, governments, and other
purchasers evaluate the full environmental costs associated with existing consumption and production
strategies, preventing a piecemeal approach. An LCA can help avoid shifting the environmental burden
from one stage to another, from one geographic area to another, and from one environmental domain or
protection target to another. For example, focusing on making improvements in one part of the life cycle
(for example, in production) may lead to even higher impacts in other parts of the same life cycle (for
example, the product use), and without LCA, the added impacts might not be considered. To take another
example, without LCA, it would be very easy to overlook the fact that a reform that reduced one type of
environmental impact associated with a product (for instance, air pollution) might increase the impact in
another part of the environment (for instance, solid waste in a landfill).
In an LCA, the use of resources, raw materials, parts and products, energy carriers, electricity, and so on,
are documented as inputs for each single step in a process. Emissions to air, water, and land, as well as
waste and other by-products, are recorded as outputs. For inputs not coming directly from raw materials
(for example, other manufactured products), their life cycle, which is often referred to as their
environmental history, has to be included in the analysis as well. For outputs, the subsequent processes
they undergo need to be included in the analysis also.
For example, in the case of a car, raw materials are first extracted from the Earth, such as mineral ores,
water, and oil. Second, raw materials are processed into finished materials, such as bauxite ore, that is
processed into aluminium, and oil that is refined into plastics. Third, the materials are manufactured or
assembled into a final product—in this case, a car. Often, this stage can be split into two parts. In the first
materials are manufactured into parts (for example, an aluminium sheet is manufactured into an
automobile body panel). The parts are then assembled into a final product (for example, the body panel,
along with the windows, engine, and the multitude of parts that are assembled into a car). Fourth is the
use stage, when a consumer has control of the product: the car, for example, will mainly consume oil
refined into gasoline and will emit pollutants into the air. Finally comes the end-of-life stage, when the
product is broken down into component materials for remanufacturing or recycling, or is
discarded. A sixth stage of distribution, when the materials and products are transported between stages,
should be also taken into account.

79
Integrated Bottom Line
The integrated bottom line (IBL) is an extension of the triple bottom line concept, which suggests that the
environment, society, and the economy are the three critical entities in any business operations. The IBL
goes a step further by suggesting that all three of them should be considered in unison in any
undertaking. All measures are combined into one balance sheet instead of three independent ones. The
environment and society are critical elements of the financial bottom line and should be considered as
such. The coinage of the term is attributable to Theo Fergusson of Sustainable Ventures. The precursor
term triple bottom line (TBL, 3BL, or people, planet, profit) succinctly describes the goal of
sustainability—the phrase was coined by John Elkington in 1994 and later expanded and explained in his
book, Cannibals with Forks: The Triple Bottom Line of 21st Century Business. The term sustainability was
first defined by the Brundtland Commission of the United Nations in 1987. The notion of this term
mandates that a company's responsibility is to the stakeholders rather than the shareholders, where
“stakeholders” refers to anyone who is influenced, either directly or indirectly, by the actions of the firm.
The goal of any business entity should be to coordinate stakeholder interests, instead of shareholder
(owner) profits.
Dimension Perspectives Focus KPIs
Economic Value Financial Financial performance ROI, Operating margin, PAT
Social Value Society Societal welfare % of profit spent for larger social
goods
Level of negative externality (if any)
is offset by positive actions
Customer Customer satisfaction Service rating
Organisational Knowledge and Employee retention
capacity innovation Flow of new business ideas
Environmental Ecosystem Environment Level of adoption of Cleaner
Value wellbeing management production: wastes and emission
throughout life cycle management
Dematerilisation strategies (Circular
Economy)
Level of Decarbonisation

The Triple Bottom Lines “People” or human capital in the triple bottom line refers to fair and beneficial
business practices toward employees, customers, community, and region in which the firm conducts its
business. An IBL enterprise seeks to benefit as many stakeholders as possible and not to exploit or
endanger any of them. The enterprise pays fair wages to its workers, maintains a safe working
environment and reasonable working hours, and does not otherwise exploit the community or its labor
force. In addition, the businesses proactively gives back to the community by offering services such as
healthcare and education. Quantifying this bottom line is often subjective and problematic. The Global
Reporting Initiatives (GRI) has developed guidelines to enable firms and nongovernmental organizations
(NGOs) alike to comparably report on the social impact of a business. The use of the word “planet” in the
triple bottom line refers to Earth's natural capital and resources. A company using an IBL approach
attempts to minimize the environmental impact by reducing its ecological footprint by, among other things,
carefully controlling the consumption of energy, reducing manufacturing waste, and careful disposing of
toxic materials in a safe and legal way. Managing all facets of the supply chain by addressing cradle-to-
grave issues is critical in firms using the IBL approach. This includes conducting a life cycle assessment
of products to ascertain the actual environmental cost from the growth and harvesting of raw materials to
the manufacturing to distribution to eventual disposal by the end user. Firms focused on IBL also
proactively avoid ecologically destructive practices such as endangering depletion of resources such as
overfishing or deforestation. Finally, “profit” in the triple bottom line is the capital earned by businesses.
Within an IBL system, the “profit” issues pertain to not just the firm, but to the society at large. Thus, in an
IBL approach, profitability is interpreted as traditional corporate accounting as well as social and
environmental impact.

80
In other words, traditional objectives of profitability, efficiency, and economic growth are judged
juxtaposed with their synchronization with biodiversity, ecological sustainability, equity, community
support, and maximized well-being for all types of stakeholders.

Carbon Negative
Carbon dioxide removal (CDR), also known as greenhouse gas removal, usually refers to human-driven
methods of removing carbon dioxide from the atmosphere and sequestering it for long periods, such that
more carbon dioxide is sequestered in the process than emitted. These methods are also known as
negative emissions technologies, as they offset emissions from dispersed sources and practices such as
energy systems, transportation, and land cover change.
CDR methods include afforestation, agricultural practices that sequester carbon in soils, bio-energy with
carbon capture and storage, ocean fertilization, enhanced weathering, and direct air capture when
combined with storage. To assess whether net negative emissions are achieved by a particular process,
comprehensive life cycle analysis of the process must be performed.
Alternatively, some sources use the term "carbon dioxide removal" to refer to any technology that
removes carbon dioxide, but may be implemented in a way that causes emissions to increase rather than
decrease over the lifecycle of the process.
The IPCC's analysis of climate change mitigation pathways that are consistent with limiting global
warming to 1.5°C found that all assessed pathways include the use of CDR to offset emissions. Using
existing CDR technologies at scales that can be safely and economically deployed, there is potential to
remove and sequester up to 10 gigatons of carbon dioxide per year, offsetting about a fifth of current
emissions.

Carbon Neutral
Carbon neutrality is achieved by reducing carbon emissions and balancing greenhouse gas emissions
through the purchase and cancellation of an equivalent amount of carbon offsets. The English village of
Ashton Hayes aspires to become England's first carbon-neutral village.
Becoming carbon neutral, sometimes known as CO2 neutral, most commonly refers to achieving net zero
greenhouse gas (GHG) emissions (including all greenhouse gases weighted according to their CO2-
equivalent impact). Carbon neutrality can be achieved by reducing and avoiding GHG emissions at the
point of origin/production, or by balancing greenhouse gas emissions through the purchase of an
equivalent amount of carbon offsets. This combination of activities allows companies, charities,
governments, nongovernmental organizations (NGOs), and individuals to reduce their net climate change
impact to nil. To emphasize reduction of GHG emissions is generally recognized as a “best practice,” and
to offset, if necessary, residual emissions.

Material efficiency
Material Efficiency (ME)—the delivery of goods and services with less material—is increasingly seen as
an important strategy for reducing GHG emissions in industry (IEA 2017a). Options to improve ME exist
at every stage in the life-cycle of materials and products. This includes: designing products which are
lighter, more optimised, last longer and with circular principles built-in; pushing manufacturing and
fabrication process to use materials and energy more efficiently and recover material wastes; increasing
the capacity, intensity of use, and lifetimes of product in use; improving the recovery of materials at end-
of-life, through improved remanufacturing, reuse and recycling processes. ME provides plentiful options
to reduce emissions, yet because interventions are dispersed across supply chains and span many
different stakeholders, this makes assessing mitigation potentials and costs more challenging. For this
reason, ME interventions have traditionally been under-represented in climate change scenario modelling
and integrated assessment models (IAMs) (Grubler et al. 2018; Allwood 2018). However, two advances in

81
the modelling of materials flows have underpinned the recent emergence of ME options being included in
climate scenario modelling.

Circular economy: It is a regenerative system in which resource input and waste, emission, and energy
leakage are minimized by slowing, closing, and narrowing energy and material loops. This can be
achieved through long-lasting design, maintenance, repair, reuse, remanufacturing, refurbishing, and
recycling. It’s also about development of new business models, which involve moving from selling goods
to offering high-quality services.

Circular economy (CE) is one effective approach to mitigate industrial GHG emissions and has been
widely promoted worldwide since AR5. From industrial point of view, CE focuses on closing the loop for
materials and energy flows by incorporating policies and strategies for more efficient energy, materials
and water consumption, while emitting minimal waste to the environment (Geng et al. 2013). Moving
away from the current linear mode of production (synthetically referred to as an “extract-produce-use-
discard” model), the circular economy promotes the design of durable goods that can be easily repaired,
with components that can be reused, remanufactured, and recycled (Wiebe et al. 2019).

Green marketing
Green marketing is the marketing of products that are presumed to be environmentally preferable to
others. Thus green marketing incorporates a broad range of activities, including product modification,
changes to the production process, sustainable packaging, as well as modifying advertising. Yet defining
green marketing is not a simple task where several meanings intersect and contradict each other; an
example of this will be the existence of varying social, environmental and retail definitions attached to this
term. Other similar terms used are environmental marketing and ecological marketing.
Green, environmental and eco-marketing are part of the new marketing approaches which do not just
refocus, adjust or enhance existing marketing thinking and practice, but seek to challenge those
approaches and provide a substantially different perspective. In more detail green, environmental and
eco-marketing belong to the group of approaches which seek to address the lack of fit between marketing
as it is currently practiced and the ecological and social realities of the wider marketing environment.
Greenwashing
Greenwashing (a compound word modelled on "whitewash"), also called "green sheen", is a form of
marketing spin in which green PR (green values) and green marketing are deceptively used to persuade
the public that an organization's products, aims and policies are environmentally friendly and therefore
‘better’; appeal to nature. Common examples present in the marketing of food products, alternative
medicine and natural medicine.
Evidence an organization is greenwashing often comes from pointing out the spending differences: when
significantly more money or time has been spent advertising being "green" (that is, operating with
consideration for the environment), than is actually spent on environmentally sound practices.[5]
Greenwashing efforts can range from changing the name or label of a product to evoke the natural
environment on a product containing harmful chemicals to multimillion-dollar marketing campaigns
portraying highly polluting energy companies as eco-friendly. Greenwashing is therefore a "mask" used to
cover-up unsustainable corporate agendas and policies. Highly public accusations of greenwashing have
contributed to the term's increasing use.
Critics of the practice suggest the rise of greenwashing, paired with ineffective regulation, contributes to
consumer skepticism of all green claims, and diminishes the power of the consumer in driving companies
toward greener solutions for manufacturing processes and business operations. Many corporate
structures use greenwashing as a way to repair public perception of their brand. The structuring of
corporate disclosure is often set up so as to maximize perceptions of legitimacy. However, a growing
body of social and environmental accounting research finds, in the absence of external monitoring and
verification, greenwashing strategies amount to corporate posturing and deception.

82
Sins of Greenwashing
According to the Home and Family Edition, 95% consumer products claiming to be green were
discovered to commit at least one of the "Sins of Greenwashing". The Seven Sins of Greenwashing are
as follows:
1. Sin of the Hidden Trade-off, committed by suggesting a product is "green" based on an
unreasonably narrow set of attributes without attention to other important environmental
issues.eg. organic garmet
2. Sin of No Proof, committed by an environmental claim that cannot be substantiated by easily
accessible supporting information or by a reliable third-party certification.
3. Sin of Vagueness, committed by every claim that is so poorly defined or broad that its real
meaning is likely to be misunderstood by the consumer.
4. Sin of Worshiping False Labels is committed when a claim, communicated either through words
or images, gives the impression of a third-party endorsement where no such endorsement exists.
5. Sin of Irrelevance, committed by making an environmental claim that may be truthful but which is
unimportant or unhelpful for consumers seeking environmentally preferable products.
6. Sin of Lesser of Two Evils, committed by claims that may be true within the product category, but
that risk distracting consumers from the greater environmental impact of the category as a whole.
Example: ITC
7. Sin of Fibbing, the least frequent Sin, is committed by making environmental claims that are
simply false. Example: Fiji water
Bioenergy
Energy derived from any form of biomass or its metabolic by-products. Some pathways rely more on
bioenergy with carbon capture and storage (BECCS), while others rely more on afforestation, which are
the two CDR methods most often included in integrated pathways.

Socially Responsible Business


A socially responsible business (SRB) is a generally for-profit venture that seeks to leverage business for
a more just and sustainable world. The objective of the SRBs involves more than just maximizing profits
for the shareholders; it is also about creating positive changes and making valuable contributions to the
stakeholders such as the local community, customers, and staff. In other words, the SRB is both profit-
oriented and socially responsible as these companies seek to make financial gains, and at the same time,
aim to improve the well-being of the community. In doing so, the businesses engage in the voluntary
initiatives with the aims of improving in various areas ranging from the social to environmental aspects of
the society.

The concept of SRB is considered to be the highest level of involvement between the company and the
community in which it operates. It holds a similar concept to Corporate Social Responsibility (CSR) in
terms of having a common goal to make positive contributions, minimizing harmful effects, and being a
force for good in society. The main difference is that the SRB goes beyond these activities and ultimately
aims to create a market space for itself. It can be achieved through forming partnerships and alliances
with the local community and collaborating with groups such as non-governmental NGOs. The SRB
prioritizes a socially responsible activity and demands to create a long-term relationship with their
community.

India’s National Guidelines on Responsible Business Conduct (NGRBC)


• Principle 1: Businesses should conduct and govern themselves with integrity, and in a manner
that is ethical, transparent, and accountable.
• Principle 2: Businesses should provide goods and services in a manner that is sustainable and
safe.

83
• Principle 3: Businesses should respect and promote the well-being of all employees, including
those in their value chains.
• Principle 4: Businesses should respect the interests of and be responsive to all its stakeholders.
• Principle 5: Businesses should respect and promote human rights.
• Principle 6: Businesses should respect and make efforts to protect and restore the environment.
• Principle 7: Businesses, when engaging in influencing public and regulatory policy, should do so
in a manner that is responsible and transparent.
• Principle 8: Businesses should promote inclusive growth and equitable development.
• Principle 9: Businesses should engage with and provide value to their consumers in a responsible
manner.
NGRBC Principles and Dimensions Responsible and Sustainable Business Model
Sr NGRBC Principles Dimensions
1 Integrity, Ethics, transparency, • Ethics, values, worldview
accountability • Regenerative and conscious capitalism
• New conception of transparency
• Greenwashing
• Accountability
• Good corporate governance
2 Safe and sustainable goods and • Green innovations
services • Responsible production
• Sustainable consumption
6 Respect, protect and restore the • Circular economy
environment • Material (including biomaterial) efficiency
• Green supply chain management
• Energy system transition
• Adoption level of cleaner production (wastes and
emission management)
• Level of negative externality offset by positive
actions

3 Well-being of employees • Well-being of all


4 Respect and responsiveness to all • Employee retention
stakeholders
5 Respect and promote Human rights
7 Responsible and transparent policy • Collaborative and constructive lobbying/advocacy
advocacy
8 Promote inclusive growth and • Societal welfare
equitable development • Inclusive development
9 Provide value to consumer • Customer satisfaction
responsibly • Market transformation

Others • Good financial performance


Note: Dimensions developed by Prof. Pramod K. Singh of IRMA, Anand, India

84
Responsible and Sustainable Business Model: Achieving Triple Bottom Line
Sr TBL Components Dimensions
1 Environmental Values • Circular economy
• Material (including biomaterial) efficiency
• Green supply chain management
• Energy system transition
• Adoption level of cleaner production (wastes and emission
management)
• Level of negative externality offset by positive actions
• Green innovations
2 Social Values • Responsible production
• Sustainable consumption
• Well-being of all
• Societal welfare
• Inclusive development
• Employee retention
3 Economic Values • Good financial performance

4 Enabling conditions • Ethics, values, worldview


• Regenerative and conscious capitalism
• Good corporate governance
• Collaborative and constructive lobbying/advocacy
• New conception of transparency
• Greenwashing
• Accountability
• Customer satisfaction
• Market transformation
Note: Dimensions developed by Prof. Pramod K. Singh of IRMA, Anand, India

85
07. The Next Phase of Business Sustainability
Andrew J. Hoffman (2018)

Business sustainability has come a long way. From the dawn of the modern environmental movement
and the establishment of environmental regulations in the 1970s, it has become a strategic concern
driven by market forces. Today, more than 90 percent of CEOs state that sustainability is important to
their company’s success, and companies develop sustainability strategies, market sustainable products
and services, create positions such as chief sustainability officer, and publish sustainability reports for
consumers, investors, activists, and the public at large.

This trend will not abate anytime soon. Surveys show that 88 percent of business school students think
that learning about social and environmental issues in business is a priority, and 67 percent want to
incorporate environmental sustainability into their future jobs. To meet this demand, the percentage of
business schools that require students to take a course dedicated to business and society increased
from 34 percent in 2001 to 79 percent in 2011, and specific academic programs on business
sustainability can now be found in 46 percent of the top 100 US master of business administration (MBA)
programs.

For all this interest, we should expect the world to become more sustainable. But problems such as
climate change, water scarcity, species extinction, and many others continue to worsen. Sustainable
business is reaching the limits of what it can accomplish in its present form. It is slowing the velocity at
which we are approaching a crisis, but we are not changing course. Instead of tinkering around the
edges of the market with new products and services, business must now transform it. That is the focus
of the next phase of business sustainability, and we can see signs that it is emerging.

The first phase of business sustainability, what we at the University of Michigan’s Erb Institute call
“enterprise integration,” is founded on a model of business responding to market shifts to increase
competitive positioning by integrating sustainability into preexisting business considerations. By contrast,
the next phase of business sustainability, what we call “market transformation,” is founded on a model of
business transforming the market. Instead of waiting for a market shift to create incentives for sustainable
practices, companies are creating those shifts to enable new forms of business sustainability.

Enterprise integration is geared toward present-day measures of success; market transformation ‘market
transformation’ is founded on a model of business transforming the market. Instead of waiting for a
market shift to create incentives for sustainable practices, companies are creating those shifts to enable
new forms of business sustainability. The first is focused on reducing unsustainability; the second is
focused on creating sustainability. The first attends to symptoms; the second attends to causes. The first
focuses primarily inward toward the health and vitality of the organization; the second expands that focus
to look outward toward the health and vitality of the market and society in which the organization
operates. The first will help future leaders get a job in today’s marketplace; the second will help them
develop a target for a lifelong career. The first is incremental, the second transformational.

Changing the way we do business is essential to addressing the challenges of environmental


degradation. The market is the most powerful institution on earth, and business is the most powerful
entity within it. Business transcends national boundaries, and it possesses resources that exceed those
of many nation-states. Business is responsible for producing the buildings we live and work in, the food
we eat, the clothes we wear, the automobiles we drive, the energy that propels them, and the next form
of mobility that will replace them. This does not mean that only business can generate solutions, but with
its unmatched powers of ideation, production, and distribution, business is best positioned to bring the
change we need at the scale we need it.

Sustainable Business 1.0: Enterprise Integration


In its first incarnation, business sustainability represents a market shift. Market pressures bring
sustainability to business attention through core management channels and functions. This began with
Nixon-era government regulation and grew to include insurance companies, investors, consumers,
suppliers, buyers, and others through the 1980s and 1990s. Such market pressures can emerge from
numerous sources: coercive drivers—from domestic and international regulations and the courts;

86
resource drivers—from suppliers, buyers, shareholders, investors, banks, and insurance companies;
market drivers—from consumers, trade associations, competitors, and consultants; and social drivers—
from nonprofit organizations, activist groups, the press, religious institutions, and academia.

While corporate social responsibility (CSR) is one response to such pressures, companies have sought
to improve competitive positioning by linking sustainability and corporate strategy. This involves
translating the issue into the core language of business management: operational efficiency, capital
acquisition, strategic direction, and market growth. In each case, the firm has an established model
that it can use to conceptualize the issue and formulate a response. In this way, sustainability becomes
much like any other business threat, where market expectations change and technological
developments advance, leaving certain industries to adapt or face demise while others rise to fill their
place.

For example, when insurance companies apply sustainability pressures on the firm, the issue becomes
one of risk management. When competitors apply such pressures, it becomes an issue of strategic
direction. When investors and banks do so, it becomes an issue of capital acquisition and cost of
capital. When suppliers and buyers do so, it becomes an issue of supply-chain logistics. When
consumers do so, it becomes an issue of market demand. Framed in such terms, much of the specific
language of sustainability recedes and is replaced by standard business logic. Therefore, companies
can remain agnostic about the science of particular issues (such as climate change) but still recognize
their importance as business concerns. The successful company can perform this translation process
and integrate sustainability into its existing structures and strategies.

Take Whirlpool, for example: It has improved appliance energy efficiency because it has watched
energy efficiency move from number 12 in consumer priorities in the 1980s to number three, just behind
cost and performance, today. Whirlpool and others expect those concerns to continue to grow. One
signal of this growth is the LOHAS consumers (Lifestyles of Health and Sustainability), a segment that
considers environmental attributes in purchasing decisions and was estimated to be a $355 billion
market in the United States in 2016 and a $546 billion market worldwide.

Another signal comes from impact investors, who consider environmental, social, and governance
(ESG) factors in their investment criteria. The sector reached $8.72 trillion of professionally managed
assets in the United States in 2016, or one-fifth of all investment under professional management. But it
is not just a specialized sector; this past May, financial advisory firms BlackRock, Vanguard, and State
Street cast votes in opposition to ExxonMobil management and called for the company to disclose its
climate change impacts.

These are all signs that the market has shifted and continues to shift. Today, consumers can buy
sustainable products, stay in sustainable hotels, eat sustainable foods, and use sustainable cleaning
products. While this greening of the market is a good thing, it is not actually solving the root problems it
was meant to address. Our world continues to become less, not more, sustainable.

Sustainable Business 2.0: Market Transformation


While business sustainability has been going mainstream, the world has witnessed unprecedented
human impacts on the natural environment that threaten the viability of life on Earth. To mark this shift,
scientists have proposed that we have left the Holocene and are now entering the Anthropocene, a new
geologic epoch that acknowledges the enormous influence of the world’s 7.5 billion people (to be nearly
10 billion by 2050) on the planet.

To measure that influence, they have identified nine “planetary boundaries” that represent “thresholds
below which humanity can safely operate and beyond which the stability of planetary-scale systems
cannot be relied upon.” These are what Lancaster University management professor Gail Whiteman
has called the “key performance indicators” (KPIs) of the planet, many of which are not doing so well.
While one (ozone depletion) is on the mend, scientists believe we have overshot the boundaries of
three: climate change, biodiversity loss, and biogeochemical flows (nitrogen and phosphorus cycles).
Further indicators are also blinking red, such as ocean acidification, freshwater use, and deforestation.
The remaining two boundaries—chemical pollution and atmospheric particle pollution—require more

87
data to assess. All of these disruptions are the result of system failures created largely by our market
institutions. They will have to be remedied by those institutions.

Fortunately, capitalism can be quite malleable. It is designed by human beings in the service of human
beings, and it can evolve to meet the changing needs of human beings. This has happened throughout
its history to address issues such as monopoly power, collusion, and price-fixing. Today’s pressing need
is sustainability—particularly to address climate change—and legislators are not the only ones who can
shift course. Many companies recognize this challenge and are pushing for new market models. In the
words of Unilever CEO Paul Polman, “We are entering a very interesting period of history where the
responsible business world is running ahead of the politicians” and taking on a broader role to “serve
society.”

The next phase of business sustainability calls for a transformation of the market, discarding such
outdated notions as treating the environment as a limitless source of materials and sink for waste,
seeing economic value as the only measure of nature’s worth, encouraging unbridled consumption, and
considering perpetual economic growth as even possible. Corporate decision makers have a key role to
play in facilitating this transition. Instead of accepting the rules of the market as given, they must change
them to incorporate the planet’s KPIs. For example, to turn around the KPI of climate change, the market
must go carbon neutral and eventually go carbon negative. We don’t yet know how to do that, but we
know that it cannot be done by one company or one product. It requires a change in the overall market.

Real sustainability is a property of a system. For example, the notion of an energy company installing
a wind farm and calling itself sustainable makes no empirical sense. A more sustainable energy
system incorporates the whole grid, encompassing generation, transmission, distribution, use, and
mobility. We can already see signals of this change happening as new energy sources, distributed
energy, demand-side management, smart appliances, and smart meters are beginning to transform
our conceptions of energy. Already, jobs in the clean energy sector have exceeded those in oil drilling.

But the energy renaissance goes further. Electric vehicles have the potential to change the grid, leveling
the electricity demand curve by charging at night and providing storage capacity during the day for
intermittent energy sources like wind and solar. Already, a Nissan Leaf automobile owner in Japan can
buy a transformer to power the house off the battery pack during a power failure. Research is under way
to scale this concept and allow consumers to rent their batteries to utilities while their car is parked.
Electric vehicles are also transforming the auto industry. Who could have predicted 20 years ago that
new entrants like Tesla would enjoy a larger market capitalization than General Motors?

Transport Sector: And as the shift to driverless cars continues, IT companies such as Apple and
Alphabet have entered the fray, shifting success factors in the auto sector from hardware to software,
and with them our conceptions of personal mobility. For example, as incumbents such as Ford Motor
seek to become mobility providers, they must learn to operate like the airline industry, where profits
increase when their cars spend minimal time idling. Given that today’s personal car is parked 95 percent
of the time, driverless cars can result in fewer cars on the road (at least in urban centers) as people
purchase mobility services rather than own cars. Fewer cars on the road means repurposing unneeded
roads, parking lots, garages, and service stations.

Systemic Corporate Strategies


As we see with the energy and transportation sectors, the potential scope of market transformation is
vast. To help flesh this out, we can conceive this sustainability revolution as proceeding from two initial
phases. First, corporations rethink their business strategies to play a stronger role in guiding the
sustainability of the systems of which they are a part. Second, the business model itself undergoes
reconceptualization. The first phase includes at least four new ways of conceiving their approach to
operations, partnerships, government engagement, and transparency.

New conceptions of operations: Market transformation calls for optimizing supply-chain logistics to
reduce risks from numerous factors such as disruptions due to increased storm severity caused by
climate change; current and future resource availability and price volatility; accelerating emissions and
concerns for public health and the environment; and the future resilience of business and civil society.

88
These risks can directly affect assets and operations, availability and costs of inputs, regulation of
sourcing and distribution, workforce availability and productivity, and stakeholder reputation. For
instance, Nestlé, Coca-Cola, Cargill, and General Mills have all faced threats to supply chains due to the
decreased availability of water, a once-plentiful resource now scarcer because of climate change and
overconsumption.

To better manage such operational systems, companies are moving away from linear models in which
items are created, used, and disposed of once they reach their end of serviceable life, and toward
circular models, where items are created, used, and then either restored or reprocessed to recover
energy or materials that can be used again. One key to this new vision of a circular economy is that it is
regenerative by design; it is organized to keep products, components, and materials at their highest
utility and value at all times.

For example, industrial and consumer products company Ricoh has concluded that by 2050, there will be
an insufficient supply of many reasonably priced raw materials to support its manufacturing needs. As a
result, the company is revising its business model using life-cycle analysis as the basis for decision
making and establishing a series of what it calls “Resource Smart Solutions” for product design and
manufacturing, reuse, collection, maintenance, and materials recovery. To change the system around it,
the company is also helping its customers reduce energy use, carbon footprint, and virgin material use
while also expanding its own opportunities for product refurbishing, recycling, and new designs. Targets
include reducing virgin resource use by 25 percent by 2020 and 87.5 percent by 2050. In adopting
circular economy thinking, Ricoh is striving to move beyond incremental efficiency goals to more
ambitious “net zero impact” business operations.

New conceptions of partnerships: Going beyond the supply chain, companies also look to novel
partnerships outside standard modes of shifting the market, including nonprofit organizations, the
government, competitors, and seemingly unrelated companies.

For example, as Ford increased its research and development in hybrid and electric drivetrains, it saw
an opportunity in how customers would live more electrified lifestyles overall. Together with Infineon,
SunPower, Whirlpool, and Eaton, Ford developed the MyEnergi Lifestyle program, exploring ways in
which hybrid electric vehicles, solar power systems, energy-efficient appliances, and home design can
be integrated to reduce the total carbon footprint. Similarly, Toyota Motor is seeking a broad array of
partnerships to achieve its goal of going “beyond zero environmental impact” by eventually eliminating
CO2 emissions from vehicle operation, manufacturing, materials production, and energy sources by
2050.

New conceptions of government engagement: Very few business schools offer courses on
collaborative and constructive lobbying. Indeed, the public perceptions of lobbying are generally
negative. But lobbying is basic to democratic politics as governments seek guidance on how to set the
rules of the market and usher reforms as needed. Forward-thinking companies are looking for ways to
participate constructively in policy formation.

For example, Intel was instrumental in calling attention to the horrors of tin, tantalum, tungsten, and gold
mining in the Democratic Republic of Congo. While the company could have simply stopped sourcing
such conflict minerals from the region, it did not want to create additional hardship for legal mining
operations. Instead, it helped create provisions in the Dodd-Frank Act that require the tracking and
disclosure of such mineral sourcing within the broader electronics industry.

This is not unusual. Companies are also working with governments to phase out heat-trapping HFC
chemicals and setting new efficiency standards on trucks. The Paris Agreement on climate change
would not have been possible without the powerful business interests that helped broker a deal. In each
of these examples, business took a responsible position in bringing about a sustainable shift in the
market through policy.

New conceptions of transparency: The only way that market transformation will be successful is
through trust, and trust can be gained only through greater transparency. The expansion of corporate

89
influence in society, particularly as it relates to government, will make some justifiably uneasy. But
robust reporting mechanisms can help allay those fears and also help protect companies from the
effects of misconduct, including legal liability and penalties. To be sure, companies are already
disclosing numerous sustainability indicators through established standards, such as the globally
recognized Global Reporting Initiative or Carbon Disclosure Project. But transparency goes further as
companies face increasing demands for data, for both internal management and external validation,
under the watchful eye of activists, investors, suppliers, buyers, employees, and customers. The
gathering and dissemination of such information can open up new awareness of supply-chain risks and
opportunities.

For example, IBM and partner companies are experimenting with blockchain technology to transform
visibility and traceability in complex, often opaque, global supply chains. In 2017, IBM piloted
supplychain blockchain with Walmart to address food safety in its global supply and distribution network
and plans to roll it out further with nine global agricultural companies. In another example, Nestlé
conducted an internal investigation of its Thai fish supply chains in 2014 and found forced labor and
brutal treatment of workers. But in a dramatic shift from standard practices of privacy and nondisclosure,
the company posted the report online, imposed new requirements on suppliers, and commissioned
outside auditors to assure compliance. This public disclosure compelled other companies that source
fish in Thailand to follow suit, shifting the competitive dynamics of supply-chain logistics.

New Ways of Doing Business


Market transformation not only compels more systemic business strategies but also challenges
traditional ways of conceiving business itself. It demands new conceptions of corporate purpose,
notions of consumption, and models and metrics of business success.

New conceptions of the corporation’s purpose: The dominant idea of the purpose of the corporation
as simply to make money for its shareholders took hold within business in the 1970s and 1980s. But the
narrow pursuit of shareholder value leads to excessively short time horizons for investment planning
and measures of success. It also leads to a focus on only the type of shareholder who is less interested
in sustainability efforts and, in the words of Cornell Law School professor Lynn Stout, is “shortsighted,
opportunistic, willing to impose external costs, and indifferent to ethics and others’ welfare.”

New ideas of corporate purpose are beginning to grow within business practice and education. For
example, benefit corporations are one type of innovation that seeks to integrate a broader array of
objectives than simply profits into its forms of organizing, governance, and legal statement of purpose.
And other companies are watching closely, sometime mimicking them.

This trend has caught on among MBA students who challenge conventional thinking around capitalism
and corporate purpose. At the Harvard Business School, an immensely popular course called
“Reexamining Capitalism” explores “the evolution, power, and limitations of our current capitalist systems”
and “how the ‘rules of the game’ by which capitalism is structured should change” to address the social
and environmental issues of our day.

New conceptions of consumption: Is “sustainable consumption” an oxymoron? The World Business


Council for Sustainable Development doesn’t think so, calling on businesses to “abandon the existing
consumption paradigm” and move toward “transformations in mainstream lifestyles and consumption
patterns.” Several businesses and activists have sought to put such an idea into practical use.

For example, Patagonia, through its Common Threads Initiative, encourages people to buy used
Patagonia products on eBay before going to the store to buy them new. Adbusters has long promoted its
“Buy Nothing Day,” what it calls a “24-hour moratorium on consumer spending” as a counterpoint to the
Black Friday spending spree that traditionally follows the holiday of Thanksgiving. The outdoor lifestyle
retailer and co-op Recreational Equipment (REI) closes its 149 stores on Black Friday as part of its
“#OptOutside” program. In 2016, Subaru, Google, Meetup, Upworthy, and competing outdoor brands
such as Burton, Keen, Yeti, and Prana chose to partner with the effort. In the end, resource use must be
reduced at the source, and that means developing new models of consumption.

90
New conceptions of business models and metrics: Market transformation requires a compelling
new business model to replace traditional ones that dominate business thinking. For example,
neoclassical economics and agency theory employ dismally simplified models of human beings as
driven primarily by selfishness, where those running the company (agents) will shirk or even steal from
the owner (principal) if they do the work and the owner gets the profits.

But behavioral economists have argued that real humans don’t behave as neoclassical economics
suggests we do, and legal scholars argue that managerial motivations are far more complex than a
simple principal/agent relationship and instead involve thousands of shareholders, executives, and
directors with more socially positive motivations. And new models have arisen, such as positive
organizational scholarship and appreciative inquiry, that move beyond standard cynical conceptions of
human behavior to understand how and why people are motivated to devote their work toward improving
the world around them and learn how to create the organizational conditions that will foster that activity.
These models are gaining increasing interest in business teaching, research, and practice as a way to
create a more committed and effective organization.

Other models are also beginning to gain recognition. Doughnut economics is a model of economic
growth that links social justice to efforts to stay within the planetary boundaries of the Anthropocene
epoch. Shared value is aimed at redefining capitalism by arguing that the competitiveness of a company
is closely tied to the health of the communities in which it is embedded.

Conscious capitalism is a model of business that serves the interests of all major stakeholders—
customers, employees, investors, communities, suppliers, and the environment. And regenerative
capitalism reimagines capitalism in terms that are self-organizing, naturally self-maintaining, and highly
adaptive to produce lasting social and economic vitality for global civilization as a whole. Each of these
models is seeking an amended form of capitalism that is sensitive to the constraints of the
Anthropocene.

Closely related to models of business behavior are the metrics used to define success, many of which
lead to unsustainable outcomes. For example, discount rates are used to capture the time value of
money—the fact that a dollar today is worth more than a dollar tomorrow. But a common discount rate of
5 percent leads to a conclusion that everything 20 years out and beyond is worthless. When gauging the
response to climate change, is that an outcome that anyone— particularly anyone with children or
grandchildren—would consider ethical? London School of Economics professor Nicholas Stern
answered no with an argument that used an unusually low discount rate when calculating the future
costs and benefits of climate change mitigation and adaptation.

Another problematic metric is gross domestic product (GDP). This measure of national economic health
fails to distinguish between financial transactions that add to the well-being of a country and those that
diminish it. Any activity in which money changes hands will register as GDP growth, even money spent
on recovery from natural disasters and pollution cleanup. To examine alternatives, former French
president Nicolas Sarkozy created a commission, headed by Nobel laureates Joseph Stiglitz and
Amartya Sen. Their 2010 report recommended a shift in economic emphasis from the production of
goods to a broader measure of overall well-being that would include measures for categories such as
health, education, security, and sustainability.

Reshaping Politics to Reshape the Market


A discussion of market transformation and the corporation’s shifting role in society cannot be complete
without a discussion of the current political and social climate and what impact it has on this agenda
going forward. The Trump administration denies the science of climate change and has embarked on an
agenda of loosening the regulatory environment to stimulate economic growth. This is a similar script to
that employed by President Ronald Reagan more than 35 years ago when he appointed Ann Gorsuch
Burford to lead the US Environmental Protection Agency, James Watt to head the US Department of the
Interior, and Rita Lavelle to run Superfund, the program for cleaning up the country’s most polluted sites.

Reagan’s appointments set about slowing or stopping environmental enforcement, but they ultimately
led to scandals and created a critical public backlash: In 1983, all three were removed from office, and

91
in subsequent years, Congress went on to strengthen numerous environmental regulations, and
environmental groups increased membership and budgets. In the words of former Sierra Club executive
director Carl Pope, Reagan “reinvented the environmental movement by his contempt for it.”

While President Trump’s approach to the environment bears similarities to Reagan’s attempts to roll
back environmental regulations and likely faces a similar backlash, there are several key differences.
First, some of the backlash this time will come from businesses that are leading on greenhouse gas
reductions and not fighting government-led environmental policies, as they did in the 1980s. Indeed,
recent surveys show that 85 percent of business executives believe that climate change is real (well
above the national average of 64 percent), and many see the associated market risks and benefits.
General Mills CEO Ken Powell was not alone when he told the Associated Press, “We think that human-
caused greenhouse gas causes climate change and climate volatility, and that’s going to stress the
agricultural supply chain.” Cargill executive director Greg Page warns of food shortages if we do not act.
Such concerns represent a strong and growing perspective within the corporate sector that we have a
problem and government inaction will only make it worse.

While those who lose in a carbon-constrained world (such as fossil fuel interests) will continue to resist
acknowledging climate change, most companies see the long-term trajectory of this issue and do not
see the current administration’s position as the long-term future. The market is shifting with or without
the US government, as other national governments as well as many US state and city governments
continue to set policies. Many companies are part of global markets and see the US withdrawal from
the Paris Agreement as ceding US leadership but not stopping the market transformation that is under
way. Some markets may slow, but some may just move to other parts of the globe, such as Germany,
India, and China, where heavy investments in renewable energy and alternative drivetrains (such as
electric and hybrid) are viewed as the future of the energy and mobility sectors.

The public is also moving in favor of sustainability. Already, public opinion polls show that an increasing
number of Americans believe climate change is real. Some even show that a majority of Republicans—
including 54 percent of self-described conservative Republicans—now believe that the world’s climate is
changing and that human beings play some role in the change. This is a marked shift from 2009, when
just 35 percent of Republicans believed that climate change was real. The truth is that many Republican
politicians, congressional aides, lobbyists, and staff believe in the science of climate change as well but
are waiting for the right political cover to voice their views.

Concern for the environment is a long-term interest of the American public, one that is more latent
than urgent and top of mind. While surveys show that it ranks low on election issue topics—number
12 in one poll, behind the economy, terrorism, foreign policy, and health care— it is also driven by
saliency, and it will awaken when threatened. That awakening can be triggered by any number of
levers. If history is any indication, smart business leadership will read these signs, anticipate the
market shift, and seek to take advantage.

References
1. John R. Ehrenfeld, Sustainability by Design: A Subversive Strategy for Transforming Our Consumer Culture, New Haven, Conn.:
Yale University Press, 2009.
2. Andrew J. Hoffman, From Heresy to Dogma: An Institutional History of Corporate Environmentalism, Stanford, Calif.:
Stanford University Press, 2001.
3. Andrew J. Hoffman, Competitive Environmental Strategy: A Guide to the Changing Business Landscape, Washington, D.C.: Island Press,
2000.
4. Andrew J. Hoffman, Getting Ahead of the Curve: Corporate Strategies That Address Climate Change, Arlington, Va.: Pew Center on
Global Climate Change, 2006.
5. Paul Crutzen, “Geology of Mankind,” Nature, 415, no. 23, 2002.
6. Will Steffen et al., “Planetary Boundaries: Guiding Human Development on a Changing Planet,” Science, 347, no. 6223, 2015.
7. John R. Ehrenfeld and Andrew J. Hoffman, Flourishing: A Frank Conversation about Sustainability, Stanford, Calif.: Stanford
University Press, 2013.
8. Terry Yosie, “Moving the Circular Economy from Concept to Business Strategy and Operations,” The Conference Board,
September 25, 2017.
9. Lynn A. Stout, “The Problem of Corporate Purpose,” Brookings Institution, 2012.
10. WBCSD, A Vision for Sustainable Consumption, Geneva, Switzerland: World Business Council for Sustainable Development, 2011.
11. Kate Raworth, Doughnut Economics: Seven Ways to Think Like a 21st-Century Economist, Chelsea, Vt: Chelsea Green Publishing,
2017.
12. Michael E. Porter and Mark R. Kramer, “Creating Shared Value,” Harvard Business Review, January-February, 2011.
13. Nicholas Stern, The Economics of Climate Change: The Stern Review, Cambridge, UK: Cambridge University Press, 2007.
14. Joseph E. Stiglitz, Amartya Sen, and Jean Paul Fitoussi, Mismeasuring Our Lives: Why GDP Doesn’t Add Up, New York: The New Press,
2010.

92
08. Circular Economy1
1. What are the most urgent changes required in the system?

For several centuries, most societies have pursued development using a linear economy model, where
the majority of resources are extracted, processed, converted to products (some of which have a very
short lifespan) and are then disposed of after use (commonly referred to, as the “take, make, waste”
process). Within this economic model, only a small percentage of materials is reused or recycled (the
exception being commodities like iron and gold). Instead, at the end of life they are considered waste and
there is often a high price, financially, socially and environmentally to dispose of this waste.

Figure 1: Material cycle

The linear economy assumes that there will always be an abundant supply of raw materials and unlimited
capacity to dispose of waste in the natural environment. However, human societies cannot continue to
operate in this way if we want to meet the demands of a growing population, preserve the health of the
planet, and ensure that future generations are able to prosper. Continuing to extract natural resources
such as minerals using this model implies an increasing environmental impact to extract ever diminishing
ore grades. The example of fossil fuel resources shows that the capacity of ecosystems to absorb
emissions is limited. Within the sustainability framework, some resources are finite and current levels of
consumption are not compatible with reaching the SDGs. An alternative is to build sustainable economies
that recognise the value of natural resources through a ‘circular economy’ (Figure 2).

The components and strategies of a circular economy model were first identified in the early 1980s and
refined in following decades (Stahel and Reday-Mulvey 1981; Ayres 1994). These earlier models referred
only to waste management – collection, separation, recycling, reuse. Today, there are many circular
economy strategies being applied by individuals, businesses and governments. These can go beyond
dealing with waste to include better product design, reduced consumption and sustainable materials
management. The common aim is to use resources in the most efficient way for the longest possible time.
The resources circulate through various processes, being reused, repaired, redesigned or

1 Reproduced from GEO 6: Global Environment Outlook 2019

93
remanufactured, which reduces the need for new raw materials and minimizes waste (Figure 2). When
faced with persistent environmental problems such as climate change, resource scarcity and biodiversity
loss, adopting resource circularity makes sense; however, society has been slow to adopt this model or
has simply failed to take the actions necessary for large-scale change.

Figure 2: Building a circular economy

Speeding up the transition to a circular economy involves a large shift in business and consumer thinking,
demanding the adoption of sustainable production and consumption processes. Fuenfschilling and
Truffner (2014) identify breaking down long-standing rigid and interdependent system structures as the
main challenge. The difficulty stems from having to enact large-scale socio-institutional change, which
may require radical new ways of thinking and adjustment to normal customs and beliefs (Potting et al.
2017). Moving from the established way of thinking involves the development of new laws and policies,
which need revised, redesigned or new business models that integrate industries and incorporate a
longer-term perspective, the internalization of the environmental and social costs of extraction, production

94
and disposal, innovative technologies, and changes in consumer use patterns. Actions that can contribute
to accelerated transformation have been outlined by the Government of the Netherlands (2016) and
include the following.
• Decreasing demand for raw materials by increasing the efficiency of raw material use in the
supply chain.
• In instances where raw materials are required, replacing fossil-based, scarce and non-sustainably
produced raw materials with sustainably produced, renewable and readily available raw
materials.
• Developing new innovative low-carbon production methods and smart product design.
• Promoting thoughtful consumption (e.g. reuse, smart design, extension of product life through
design and repair, use of secondary and recycled materials, sharing economy).

Circular economy strategies have also been developed by Germany, Finland, Denmark and Slovenia.
France, Italy and Spain have their road maps developed as well. The circular economy promotes a
production and consumption model that includes restoration and regeneration where possible (Ellen
MacArthur Foundation 2015; Smol, Kulczycka and Avdiushchenko 2017). It ensures that the worth of
products, materials, chemicals and resources is maintained in the economy at their highest utility and
value for as long as possible (European Commission 2015; Stahel 2016). The circular economy,
therefore, means reducing waste during production, ensuring asset recovery including waste utilization,
and developing obsolescence prevention pathways in product and urban system designs through
sustainable materials management. It also means ensuring product and service delivery with energy and
materials from renewable sources, while changing business models to match these objectives (Ghisellini,
Cialani and Ulgiati 2015; Rizos, Tuokko and Behrens 2017).

The circular economy preserves raw materials, thereby decoupling economic growth from the use of
resources and its associated environmental externalities, including carbon emissions. However, in some
cases the appearance of growth decoupling in one sector or territory can mask a continued environmental
and social impact somewhere else (details in Ward et al. 2017). Ward et al. (2017) cite substituting one
non- renewable resource for another (e.g. the cleaner energy systems that replace fossil fuels still require
non-renewable resources) and shifting the cost somewhere else (e.g. importing resource- intensive
consumer goods from developing countries).

2. What are the elements of the system that the policies seek to address?

Policies that support the transition to circularity are being developed and implemented in many places
and involve a range of different approaches. Early examples include the German Closed Substance
Cycle and Waste Management Act introduced in 1996 to recover materials from municipal and
production waste (Germany, Federal Ministry for the Environment, Nature Conservation and Nuclear
Safety [BMU] 2011) and the Japanese recycling initiative Basic Law for Establishing a Recycling-Based
Society (Environment Agency Japan 2000). These actions are examples of what has become known as
the 3Rs of reduce, reuse and recycle, and are the foundation of green manufacturing and consumption
(Jawahir and Bradley 2016). However, in the last decade the focus has expanded from ‘green’ to
sustainable manufacturing – for example, the 6Rs of manufacturing, which in addition to reduce, reuse
and recycle, include recover (for a subsequent life cycle), redesign (the next generation of products) and
remanufacture (meaning restoration to an ‘as new’ form) (Jawahir and Bradley 2016).

China adopted the circular economy as a development strategy in 2002, and this was given legal effect in
2009 through the Circular Economy Promotion Law (China, National People’s Congress 2008). The
European Commission released a ‘Roadmap to a Resource Efficient Europe’ in 2011, which was
replaced in 2015 by ‘Closing the Loop: An EU Action Plan for the Circular Economy’ (McDowall et al.
2017). Both Europe and China were following earlier research and policy work in the United States of
America, Japan and Europe that focused on waste management.

Table 1: Examples of the policy focus to achieve key elements of the circular economy (CE)

95
Key elements Policy examples Result examples
of CE
Design for the EU Ecodesign Directive – ensures energy It is estimated that the Ecodesign Directive
future efficiency of products, such as household will deliver a 16 per cent reduction in the
appliances, by setting minimum efficiency primary energy consumption of 35 product
requirements (EU 2009). groups compared with the consumption of
these products in 2010. For example, the
energy efficiency of televisions, under the
Ecodesign scenario it is predicted to
improve by a factor of 25 (measured from
1990) by 2030 (European Commission
2017).
Market-based Taxes on virgin materials such as sand, The United Kingdom of Great Britain and
instruments – gravel and rock used in the construction Northern Ireland introduced a tax on
green taxation industry have been introduced by 16 EU aggregates in 2002. Since the introduction
states. of the tax, primary aggregates use has
reduced by approximately 40 per cent per
unit of construction (Ettlinger 2017).
Incorporate Republic of Korea has some of the world’s Streaming music reduces resource use
digital fastest internet speeds, with connections to and costs 80 per cent less than the cost of
technology more than 90 per cent of the population. producing and distributing of compact
The government has provided economic disks (CDs) (Lacy 2015). The Republic of
support for broadband infrastructure Korea was the sixth top music market in
development, subsidies to ensure 2017 and has the largest number of paid
connectivity, and measures to stimulate music subscribers (International
information technology literacy (Falch and Federation of Phonographic Industry [IFPI]
Henten 2018). 2018).
Collaborate In Sydney, Australia, the city council GoGet is an Australian car-sharing
introduced policies to promote car sharing, company, operating in large cities.
including the provision of designated car- Members have access to a range of
share parking spaces; and online listing of vehicles including cars and vans (GoGet
private vehicles participating in peer-to- https://www.goget.com.au).
peer sharing schemes (City of Sydney
2016).
Use waste as In 1997, Denmark introduced legislation The Kalundborg Symbiosis in Denmark is
a resource that banned sending waste that could be a network of businesses that was the first
recycled or incinerated to landfill. In 2015, industry group to fully develop industrial
a new law was introduced, the symbiosis. The collaboration includes a
Environmental Technology Development coal-fired power plant, fish farming,
and Demonstration Programme (MUDP). fertilizer production and a host of other
This includes manufacturing and industrial operations
a subsidy scheme, innovation partnerships (Kalundborg Symbiosis 2018).
and international cooperation to find
resource-efficient solutions to
environmental problems (Denmark,
Ministry of Environment and Food n.d.).
Rethink the New business models that utilize The China Construction Bank Corporation
business technologies are emerging, such as (CCB) is using the IBM Blockchain platform
model blockchain. Estonia, for example, has to improve procedures for the sale of its
established an e-residency scheme to insurance products.
encourage entrepreneurs. E-residency
provides anyone with
a digital ID that allows them to access
Estonia’s e-services for online business
development and management from
anywhere in the world.
Preserve and The right to repair – the EU is preparing Inrego, a Swedish firm, is refurbishing

96
Key elements Policy examples Result examples
of CE
extend legislation making it mandatory for electronic equipment such as laptops,
existing companies to provide spare parts and personal computers, monitors and phones
products diagnostic tools that would make it cheaper (European Remanufacturing Network
and easier to repair products (European 2018).
Parliament, Committee on the Internal
Market and Consumer Protection 2017)
Prioritize Norwegian policies to support battery In Norway, incentive programmes to
regenerative electric vehicles (BEV): zero annual road encourage use of BEVs began in the early
resources tax (2018); 40 per cent reduced company 1990s. Norway currently leads the world
car tax (2018); 50 per cent price reduction with 21 per cent BEV market share (cf.
on ferries (2018); zero re-registration tax Australia, where there is limited incentive
for used zero-emission cars (2018); free and BEVs have 0.2 per cent of the market
municipal parking in many cities (Norsk (ClimateWorks Australia 2018).
elbilforening 2018).

3. What has been done to date and how effective have these measures been?
Many governments have introduced policies and regulations that address aspects of the circular
economy. Policies supporting the circular economy can focus on one or more elements of the ‘take, make
and waste’ process. While many policies have tended to address waste through recycling and resource
recovery, there are significant gains to be made at the earliest stages of product design and manufacture.
For example, products can be designed using eco-design principles, to use less material and last longer.
They can be refurbished or repaired and made of non-toxic materials that are simple to recycle.

Policies that encourage eco-design also need to consider the potential adverse health, gender and
developmental impacts of poorly planned policies (e.g. toxic exposures for women and children from
recycling electronic waste). About 15 million people are involved in informal waste recycling of plastics,
glass, metals and paper where these activities are a risk both to the environment and to the people
performing the tasks (Yang et al. 2018). Individuals performing resource recovery, especially e-waste
pickers in developing countries, risk considerable occupational and environmental health threats (Velis
2017). Women and children are among the vulnerable groups working in this informal sector who face
exposure to hazardous chemicals and heavy metals (Heacock et al. 2016), with few to no measures for
prevention or treatment (Han et al. 2018).

Policies can also support the move from managing waste to more environmentally sustainable outcomes,
by focusing on behaviour change. These policies, which are often developed from grass-roots initiatives,
aim to limit the amount of waste produced and increase material recovery (Silva et al. 2017).

Europe has established policies for implementing the circular economy, while in other areas this has
happened at a national or subnational level. There have also been some international policy initiatives
that align with, or promote, a circular economy approach, especially with regard to waste minimization
(e.g. the Basel and Stockholm conventions). The new approach of green (or sustainable) chemistry is
working to develop alternative solutions aimed at eliminating or at least significantly reducing hazardous
chemicals and eventually their presence in the environment (Weber, Lissner and Fantke 2016). One of
the challenges in relation to chemicals and the circular economy is increasing recycling and reuse, while
making sure consumers are not at risk from exposure to substances of concern that may be present in
products and passed on through waste (European Commission 2015). For some chemicals and toxic
metals, such as persistent organic pollutants (POPs) and mercury, final disposal may be a better option
than recycling and reuse.

97
09 Industrial system transition for a dematerialised and decarbonised world1

Pramod K. Singh2 and Harpalsinh Chudasama3

Abstract

The concept of industrial system transition, introduced in the IPCC special report on Global Warming
of 1.5°C remains poorly conceptualised. We, in this paper, deepen the conceptualisation of industrial
system transition and chart its pathways using fuzzy cognitive maps, which is used to model complex
and hard-to-model systems. The industrial system transition involves interactions between
dematerialisation and decarbonisation goals, along with the enabling factors of governance and
systemic corporate strategies. Material efficiency and circular economy are the main dematerialisation
propellers, while innovations in energy sources and energy efficiency are the key decarbonisation
choices. Our respondents of fuzzy cognitive maps-based surveys were 106 experts and practitioners
from industries, with about 55% of respondents representing global fortune 1000 companies. Fuzzy
cognitive map-based simulations reveal that governance, policies and regulations in terms of better
transnational governance, technology push, market-pull, carbon price and carbon market, and
technology transfer from developed to developing countries could play an extremely crucial role in
attaining decarbonisation and dematerialisation goals. Our findings also confirm that enabling corporate
strategies in terms of regenerative and conscious capitalism and market transformation plays a
significant role in achieving the goals of decarbonisation and dematerialisation, albeit to a lesser degree
than governance, policies, and regulations. However, the stellar performance of industrial system
transition to dematerialisation and decarbonisation requires the combined measures of (i)
dematerialisation and decarbonisation, (ii) governance, policies, and regulations (technology push,
market-pull, carbon price and carbon market, and technology transfer from developed to developing
countries); and (iii) systemic corporate strategies (regenerative and conscious capitalism, and market
transformation strategy). Since large industries are in the hands of transnational corporations,
appropriate transnational governance strategies are sine-qua-non for achieving the goals of
dematerialisation and decarbonisation. Several transnational governance networks have engaged in
public-private co-governance in the decarbonisation space dominated by larger players. The advent of
polycentric governance provides new possibilities for trans-local governance opportunities where large
numbers of small and medium industries can participate in advancing at least decarbonisation goals;
however, such networks require supports from the governments.

Keywords: Decarbonization; Materials efficiency; Circular economy; Environmental governance;


Environmental strategy; Sustainable production.

1. INTRODUCTION

Industrialisation, a vital development driver, is also a symbol of modernity. Industrialisation contributes


to human well-being by generating employment, economic growth, reducing poverty and inequalities,
and social inclusiveness (UNIDO, 2013). As of 2016, energy use contributed to 69.7% of the total
greenhouse gas (GHG) emissions, while industry and waste contributed 8.3% of the total GHG
emissions (Ritchie and Roser, 2019). Thus industrialisation is responsible for depleting planetary health.

1
Under review with Global Environmental Change, reproduced for class discussion only.
2
Professor, Institute of Rural Management Anand, Email: pramod@irma.ac.in
3 Research Associate, Institute of Rural Management Anand, Email: harpalsinh@irma.ac.in
1

98
1.1. Conceptualising Industrial System Transition

The IPCC special report on Global Warming of 1.5°C introduced the concept of industrial system
transition. As per this report, industrial system transition involves carbon dioxide capture utilisation and
storage (CCUS), energy efficiency, bio-based and circularity, shifting to low- or zero-emission power
generation, such as renewables and hydrogen, electrifying transport and developing green
infrastructure, and improving energy efficiency by smart urban planning (de Coninck et al., 2018). The
International Energy Agency (IEA) identified six partly overlapping, and interdependent mitigation
options in the industrial sector: (i) demand management, (ii) materials efficiency, (iii) circular economy
and industrial waste, (iv) energy efficiency, (v) electrification and fuel switching, and (vi) CCUS,
feedstock, and biogenic carbon fuel switching (IEA, 2019; 2017).

Material efficiency—the delivery of goods and services with less material (IEA, 2019; Worrell et al.,
2016), circular economy along with demand management of raw materials (Garmulewicz et al., 2018),
green innovations (Marchi, 2012), and green supply chain management (Abhishek and Divyashree,
2019) lead to dematerialisation measures. Material demand has been cited as the main explanation for
driving energy consumption and industrial emissions (IEA, 2019). Besides, material efficiency is
increasingly seen as an essential strategy for reducing industry-related GHG emissions (Material
Economics 2019; IEA, 2017). Thus both dematerialisation and decarbonisation are needed for industrial
system transition.

The sustainability transitions involve multiple, co-evolving elements including ‘technologies, markets,
infrastructures, policies, industry structures, and supply and distribution chains’ (Kohler et. al., 2019),
which organise themselves into stable regimes often understood as consisting of technologies,
institutions and actors. Hence the role of governance, policies, and regulation cannot be over-ruled.

Governments have been increasingly regulating the materials use and their disposal worldwide, which
is likely to result in the restricted use of many materials and increased recycling (Rankin, 2014).
Furthermore, a policy drive in terms of technological transition is needed to induce material efficiency
(Allwood et al., 2019; Rankin, 2014), circular economy (Garmulewicz et al., 2018), and demand
management of raw materials. Similarly, a market-pull in terms of high and rising consumer demand
for material efficiency (Rizos et al., 2016) and raw materials’ management (Olatunji et al., 2019) is
likely to induce manufacturers to invest in decarbonisation. Market-pull, powered by end-customers,
will encourage the adoption of new technologies, energy-source developments (Allwood et al., 2019),
and energy efficiency (Kang and Lee, 2016) when there is a direct correlation with increased revenue
and returns. Additionally, carbon pricing has been observed to be a vital tool for low-cost and low-
carbon mitigation strategies (Ryan et al., 2011).

Systemic corporate strategies such as regenerative and conscious capitalism, market transformation,
collaborative and constructive lobbying, and a new conception of transparency are shaping the next
phase of business sustainability (Hoffman, 2018). However, greenwashing, the deceptive corporate
strategies often exert profound adverse effects on consumer and investor confidence and goes against
business sustainability (De Jong et al., 2017; Newell et al., 2015; Delmas and Burbano, 2011).

Thus industrial system transition involves dematerialisation and decarbonisation which are facilitated
by governance, policy and regulation, and systemic corporate strategies. Dematerialisation involves
materials efficiency (Worrell et al., 2016), circular economy (Garmulewicz et al., 2018), demand
management of raw materials, green innovations (Marchi, 2012), and green supply chain management
(Abhishek and Divyashree, 2019). On the other hand, decarbonisation comprises innovations in energy
sources and energy efficiency (Kang and Lee, 2016). The role of technology push, market-pull, carbon
price and carbon market (Ryan et al., 2011), and transnational governance (Andonova et al., 2009) is
2

99
crucial for industrial system transition. Systemic corporate strategies of regenerative and conscious
capitalism, market transformation strategy, collaborative and constructive lobbying, and a new
conception of transparency (Hoffman, 2018) could be other enabling factors of industrial system
transition.

1.2. The State of Knowledge in Industrial System Transition-Related Areas

In the recent wake of the COVID-19 pandemic, with significant declines in demand for fuels, oil, coal,
gas, and nuclear power, there is concern that the rebound from the crisis will harm the environment
(IEA, 2020). This requires a shift to greener, more efficient, and renewable electricity networks
(Hosseini, 2020) and decarbonisation. The former involves limiting carbon emissions in energy-
intensive industries like material extraction and processing (Hildingsson et al., 2019). The latter entails
a reduction in material intensity, recycling, and reusing of secondary material leading to a reduction in
resource use (Petrides et al., 2018). Adopting these mechanisms could enhance planetary health and
human wellbeing (Waddock, 2020).

Options to improve material efficiency exist at every stage in the life-cycle of materials and products
including designing products which are lighter, more optimised, last longer and with circular
principles built-in; pushing manufacturing and fabrication process to use materials and energy more
efficiently and recover material wastes; increasing the capacity, intensity of use, and lifetimes of
product in use; improving the recovery of materials at end-of-life, through improved remanufacturing
reuse and recycling processes (IEA, 2019).

The circularity model of a circular economy includes refuse, rethink, reduce, reuse, repair, refurbish,
remanufacture, repurpose, recycle, and recover energy, meaning less energy and resource demand
(Potting et al., 2018). Circular economy tools and policies at the micro-level include cleaner production,
eco-design, environmental labelling, process synthesis, and green procurement. The 1.5°C Low Energy
Demand (LED) scenario produced by Grubler et al. (2018) considers a 20% decline in global material
output from today, by 2050. Industrial decarbonisation is possible on the mid-century horizon but
requires regionally and sectorally specific term policy strategies that consider whole innovation and
supply chains (Bataille et al., 2018; Åhman et al., 2017; Grubb et al., 2017).

Initiatives to incorporate material efficiency, circular economy, and demand management of raw
materials, along with innovation in energy sources and energy efficiency to promote sustainability at
various levels of the industrial system, must be assisted by appropriate government policies (Rankin,
2014). Incentivising low-carbon technologies and innovations in addition to internalising the cost of
GHG emissions in consumer and producer investment decisions are likely to promote energy source
innovation and energy efficiency while mitigating carbon emissions cost-effectively (Boyce, 2018;
Schmalensee and Stavins, 2017). Likewise, technology transfer from developed to developing countries
will facilitate the dematerialisation and decarbonisation measures (Hansen et al., 2019). Technology
transfer from developed to developing countries is likely to improve material efficiency (Heshmati,
2018; Allwood et al., 2013) and expedite green innovations measures for dematerialisation (Hansen et
al., 2019; Lema and Lema, 2012) while improving energy efficiency (Wei et al., 2016) in the early
stages of modern industrialisation and ensuring decarbonisation (Hansen et al., 2019).

There are several barriers hindering corporations, industries, and governments from leveraging
opportunities to ensure a sustainable industrial transition. The extant literature cites the economic health
of industries (Lereboullet et al., 2013), limited regulations (Hurlimann et al., 2018), the lack of
investments (Fischedick et al., 2014;), the lack of incentives (Fischedick et al., 2014), unavailability
and access to infrastructure, information, and market instruments such as credit and insurance
(Fischedick et al., 2014), market risks (Jacobsson and Bergek, 2011), the geographical location of
3

100
industries (Holland and Smit, 2014), and internal processes of the company (Long et al., 2018), as
common hindrances. The effectiveness of decarbonisation and dematerialisation initiatives, however,
depends on the environment under which these industries operate. Therefore, a key step towards a
structural and sustainable industrial transformation is to lighten the obstacles preventing its adoption.

Demand for decarbonisation and dematerialisation strategies is driven by the industry’s desire to
participate in sustainable production together with the policy and governance that determines the
ecosystem in which industrial processes occur. Greater co-operation could help information, market
development, and technology penetration (Allwood et al., 2019; Goodman et al., 2017). In literature,
theories like circular economy (Potting et al., 2018), regenerative economy (Fullerton, 2015), and
conscious capitalism (Mackey and Sisodia, 2014) are frequently proposed as alternative frameworks
for systemic transitions in business strategies.

Regenerative and conscious capitalism is a way of rethinking capitalism and industries. These principles
will help industries create a self-organising, self-maintaining, highly adaptive, regenerative form of
capitalism that generates lasting social and economic vitality and human well-being (Fullerton, 2015).
Simultaneously, market transformation strategies are important while addressing the challenges of
environmental degradation and are likely to change the way business is conducted (Hoffman, 2018).
Here ‘market transformation’ portrays a business model that transforms the market (Hoffman, 2018).

Transparency in material production, trade, and consumption is another step that industries can take to
ensure a sustainable transition. Transparency helps stakeholders simplify supposedly complex supply
chains (Gardner et al., 2019) and, subsequently, help identify and assess dematerialisation and
decarbonisation opportunities, and also reverse existing unsustainable practices. The need for
transparency has led to increased pressure from industry stakeholders, investors, governments, and
consumers to publish their environmental performance and eliminate greenwashing (Wu et al., 2020;
Marquis et al., 2016; Kim and Lyon, 2015). When companies engaged in greenwashing, the consumers
and investors are often reluctant to reward them, notwithstanding their environment-friendly
performance (De Jong et al., 2017; Newell et al., 2015; Delmas and Burbano, 2011). This, in effect,
discourages companies from engaging in environmentally detrimental behaviour, causing negative
externalities and yielding a negative effect on bottom-lines (ibid).

1.3. Knowledge Gap

The extant literature is scarce on what makes materials efficiency, circular economy, innovations in the
energy sources, and energy efficiency work, which is incredibly crucial for industrial system transition
and achieving the Sustainable Development Goal 12—the sustainable production and consumption.
Even the integrated assessment models (IAMs) in their climate change scenario modellings
underrepresent the materials efficiency and circular economy. Studies (IEA, 2019; Material Economics,
2019) propose six overarching planning and strategies for industrial transition pathways: (i) Policies to
encourage material efficiency and high-quality circularity; (ii) supply pushing R&D and early
commercialisation along with demand and pull lead/niche markets to help emerging technologies cross
the valley of death; (iii) carbon pricing or regulations with competitiveness provisions to trigger
innovation and systemic carbon efficiency; (iv) long-term, low-cost finance mechanisms to enable
investment and reduce risk; (v) infrastructure planning and building; (vi) a shared global R&D effort
towards achieving low to zero-carbon technologies. However, the extent of facilitation by these factors
is unknown and this study tries to fill the gap.

101
1.4. Why FCM?

Considering the knowledge gap cited in the preceding section and to model complex interactions of
industrial system transition, there is a need for a right approach and methodology. Majority of the extant
literature relating to ‘dematerialisation’, ‘decarbonisation’, ‘material efficiency’, ‘circular economy’,
‘demand management’, ‘green innovation’, ‘green supply chain management’, ‘energy efficiency’,
‘transnational governance’, ‘technology push’, ‘market pull’, ‘carbon pricing’, ‘carbon market’, and
‘conscious capitalism’ have adopted either quantitative or qualitative research methods. These include
life cycle assessment, material flow analysis, structural equation modeling, analytic hierarchy process,
multi-criteria decision making, system dynamics, social network analysis, fuzzy set theory, and multi
and/or hybrid modelling approaches. Some studies have adopted a mixed approach including both
quantitative and qualitative methods. Substantial studies have adopted a conceptual approach, either
following a systematic review of literature or descriptive and/or exploratory and explanatory research
study including a case study based research. There is hardly any study, which used the fuzzy cognitive
map (FCM), which is used to model complex and hard-to-model systems and extent of relative
contributions of different factors. The industry as a whole is a very complex web of sectors, sub-sectors,
and inter-sectoral interactions and dependence, with associated mitigation opportunities and co-benefits
(OECD, 2019). Hence we adopted FCM-based simulations, which proves to be handy in modelling
such a complex system. The FCMs capture stakeholders’ knowledge, perceptions, and beliefs for
evidence-based decision-making in environmental planning and management (Singh and Chudasama,
2017a,b; Solana-Gutiérrez et al., 2017) and reveal the role of hidden and important feedbacks in the
system (Özesmi and Özesmi 2004).

The rest of the paper is structured as follows: In section 2, we discuss FCM methods adopted for
knowledge elicitation from stakeholders and FCM-based simulations. Section 3 presents findings and
discussions on the factors required for the transformation of the industrial system to sustainable
production. The discussion and conclusions sections describe the necessary governance, policy, and
regulatory frameworks required for the industrial sector transformation.

2. METHODS

We adopted FCMs to understand stakeholders’ understanding of the transitions required in the industrial
systems for sustainable production. FCM is a quasi-quantitative research method used to understand the
functioning of a dynamic and complex system. The FCMs allows the study participants to communicate
their perception and knowledge through causal relationships between various factors called concepts
(Özesmi and Özesmi, 2004; Kosko, 1986).

2.1. Knowledge Elicitation

An FCM-based study adopts two approaches of concept design: ‘open-concept design’ and ‘pre-
designed concept’ (Singh and Chudasama, 2020). The pre-designed concept often use interviews (Ziv
et al., 2018). We adopted a ‘pre-designed concept’ approach using a literature review (Morone et al.,
2019). We drew up a list of 62 statements representing causal relationships between the 26 concepts
based on our conceptualisation of industrial system transition and the literature review. We provided
the statements representing causal relationships and definitions of the concepts to elicit weights from
these causal linkages to the respondents. We sent the instruments to 1360 respondents, comprising
global fortune 1000 companies (except financial sector), non-fortune 1000 companies representing
aerospace and defense, automobile, aviation industry, building products and equipment, cement,
chemicals, chemicals, electrical equipment, energy, engineering and construction, heavy equipment,
mining, and steel sector. We also sent the instruments to 44 authors who have published in relevant
fields since 2014. We received 106 full and 11 partial responses, which were excluded. Participants
5

102
assigned values to each question on a scale of 1 to 10, showing the intensity of the relationships between
various factors called concepts (Özesmi and Özesmi, 2004). The values ranging within 1–2 signified
very low, 3–4 low, 5–6 moderate, 7–8 high, and 9–10 very high strengths. An arithmetic mean of the
weight values was calculated to comprehend the cumulative perception of multiple study participants
(Singh and Chudasama, 2020; Singh et al., 2019; Singh and Chudasama, 2017a,b; Özesmi and Özesmi,
2004).

2.2. Constructing Adjacency Weight Matrix

By listing them in vertical and horizontal axes to form a square adjacency weight matrix, the concepts
were recorded on an excel sheet (Singh and Chudasama, 2017a,b). The values given by the participants
were normalised between –1 and +1 before coding them into an adjacency matrix (Singh and
Chudasama, 2017a,b; Özesmi and Özesmi, 2004). The weight values were coded into the adjacency
matrix for each link between the concepts (Özesmi and Özesmi, 2004).

2.3. FCM-based Simulations

The FCMs can be represented using four functions: M = (C, W, A, f): (i) C is the total number of concepts
in the map, (ii) W : (Ci, Cj) → wij represents the adjacency matrix of the map, (iii) WM×M, A : (Ci) → A(t)i
computes the activation degree of each concept Ci at the discrete-time step t (t = 1, 2, . . . , T), and (iv)
f (x) is the transformation function (Nápoles et al., 2016). Simulation results are generated by
multiplying the initial-state vector value of a concept with the adjacency matrix (Singh and Chudasama,
2020; Singh et al., 2019; Singh and Chudasama, 2017a,b; Nápoles et al., 2016). The activation rule and
the transformation function decides outcome of the simulations.

The activation rule computes the activation vector using the initial-state vector value of the concepts
(ibid). We used the activation rule introduced by Stylios et al. (1997) because of its memory capabilities
(equation 1).

𝑛
(𝑡+1) (𝑡) (𝑡)
𝐴𝑖 = 𝑓 (𝐴𝑖 + ∑ 𝑤𝑖𝑗 × 𝐴𝑗 ) , 𝑖 = 𝑗 (1)
𝑗=1

Where, Ai(t+1) is the vector value of concept Ci at the simulation step t+1; Ai(t) is the vector value of
concept Ci at the simulation step t; n is the total number of concepts and wij is the weight of the link
between concept Ci and concept Cj; Aj(t) is the vector value of concept Cj at the simulation step t; 𝑓 is
the type transformation function deployed during the simulation process.

Considering the state vector values of concepts in the range of (–1,+1) we used hyperbolic tangent
transformation function (Singh and Chudasama, 2017a,b.) as shown in equation 2.

𝑒 𝜆𝑥 − 𝑒 −𝜆𝑥
𝑓 (𝑥 ) = , 𝑎 ∈ R+ (2)
𝑒 𝜆𝑥 + 𝑒 −𝜆𝑥

(𝑡)
Here 𝑥 is the vector value 𝐴𝑖 of a concept for a given iteration; e is a residual that describes the
minimum error difference among the subsequent concepts; 𝛌 determines the steepness of the slope; and
R+ is used to keep the activation value of concepts in the (–1, +1) interval (Carvalho, 2013).

103
Each concept has a vector value that varies from |0| to |1|, meaning ‘non-activated’ or ‘activated’,
respectively (Singh and Chudasama, 2020; Singh and Chudasama, 2017a,b). ‘Activated’ concepts
influence other concepts following the weighted causal relationships of the adjacency matrix (ibid). The
FCM converges to a steady-state at the end of the simulation process (Singh and Chudasama, 2020;
Singh et al., 2019; Singh and Chudasama, 2017a,b; Solana-Gutiérrez et al., 2017; Nápoles et al., 2016;
Carvalho, 2013). Since our FCM model has multiple feedback loops, repeated activation of a concept
sets-in (Singh and Chudasama, 2020; Singh and Chudasama, 2017a,b; Nápoles et al., 2016). We
developed five scenarios for FCM-based simulations to identify the critical factors responsible for
dematerialization and decarbonisation. Concepts for simulation analysis were selected based on their
out-degree centrality in each category of concepts.

3. RESULTS AND DISCUSSION

The industrial system transition is characterised by interactions between multiple concepts along with
their causal interactions. These causal linkages across various concepts across dematerialisation,
decarbonisation, governance, and systemic corporate strategies are illustrated in Fig. 1.

3.1. Dematerialisation and Decarbonisation: Transition of Industrial Systems

Decarbonisation and dematerialisation strategies can be considered as key propellers for the industrial
system transition. The results of FCM-based simulation reveal that only technical measures of material
(including biomaterial) efficiency, circular economy, innovations in the energy sources (CDR
techniques, the role of hydrogen in the industry, etc.), and energy efficiency can help achieve a moderate
level of decarbonisation and dematerialisation goals (Fig. 2, Table 1: Scenario 1).

Industrial system transitions include shifts around and within current value chains, new sectoral
couplings, and include power, hydrogen, and other infrastructure enabling. Innovations in energy
sources and energy efficiency are the key propellers of decarbonisation (Fig. 1). New technology and
innovation in energy sources, particularly for the manufacturing sector, help to ensure the
decarbonisation (Sun et al., 2019).

Recent modelling suggests that per capita material stocks level-off in developed countries and decouple
from GDP. For instance, Bleischwitz et al. (2018) confirmed the occurrence of a saturation effect for
four materials (steel, cement, aluminium, and copper) in five industrialised countries (Germany, Japan,
UK, USA, and China). However, high growth in material supply may still drive global demand for new
products in the coming years for developing countries which are still far from saturation levels.
Although sustainable consumption at the consumers level can play a huge role in determining
industries’ willingness to integrate sustainability into their existing businesses (Allwood et al., 2019),
consumption patterns remain embedded in economic, cultural, social, and political systems (Worrell et
al., 2016). We argue that sustainable production could lead to sustainable consumption as promotion
campaigns play a crucial role in modifying consumption behaviour.

104
Fig. 1: Interacting factors of industrial system transition

105
Fig. 2: Pathways For Industrial System Transition

3.2. Governance, Policies, and Regulations: the Great Enablers

FCM-based simulation results suggest that dematerialisation and decarbonisation measures, along with
better governance, policies and regulations (effective transnational governance, technology push,
market-pull, carbon price and carbon market, and technology transfer from developed to developing
countries) strongly impact the achievement of decarbonisation and dematerialisation goals (Fig. 2,
Table 1: Scenario 2).

Globally, governments have resorted to conventional measures such as banning and taxing, along with
legislation and environmental agreements to promote material efficiency (Allwood et al., 2011). By
asserting the creation and dissemination of renewables and zero-emissions technologies and
withdrawing fossil-based value propositions from the market, governments can transition towards deep
decarbonisation (Rockström et al., 2017). However, government policies should be founded on (a)
polluters pay principle, (b) common but differentiated responsibility, (c) development, deployment and
transfer of technology, and (d) easing the conflicts between current free trade regimes and motivated
industrial policies (Åhman et al., 2017). A policy mix should thereby focus on regulations—standards
and labelling (Schwarz et al., 2019; Tanaka, 2011), material efficiency (CIWMB, 2003), banning the
use of materials (Mosquera, 2019), mandating technologies and targets (Tanaka, 2011); economic
instruments—carbon pricing and markets (Boyce, 2018), green taxes and carbon pricing (Ryan et al.,
2011), preferential loans, and subsidies (Taylor, 2008); voluntary actions agreements (UNEP, 2018);
extended producer responsibilities (Xiang and Ming, 2011); information programmes—monitoring and
evaluation (UNEP, 2018; Söderholm and Tilton, 2012), partnerships, research, and development
(Bataille et al., 2018); and government provisioning of services—government procurements (Ghisetti,
2017), technology push and market-pull (Hansen et al., 2019; Fischedick et al., 2014; Taylor, 2008).

106
Table 1: Industrial system transition scenarios

Scenario 1 Scenario 2 Scenario 3 Scenario 4 Scenario 5


Concepts
IV FV IV FV IV FV IV FV IV FV
C1: Materials efficiency 1 0.902 1 1.000 1 0.902 1 1.000 1 1.000
C2: Circular economy 1 0.131 1 0.997 1 0.131 1 0.997 1 0.997
C3: Demand management of raw materials 0 0.877 0 0.997 0 0.877 0 0.997 0 0.997
C4: Green innovations 0 0 0 0.909 0 0 0 0.909 0 0.909
C5: Green supply chain management 0 0 0 0.895 0 0 0 0.895 0 0.895
C6: Other mitigation and adaptation interventions 0 0 0 0 0 0 0 0 0 0
C7: Dematerialisation 0 0.972 0 1.000 0 0.996 0 1.000 0 0.999
C8: Decarbonisation 0 0.965 0 1.000 0 0.989 0 1.000 0 0.999
C9: Innovations in energy sources 1 0.131 1 0.998 1 0.131 1 0.998 1 0.998
C10: Energy efficiency 1 0.613 1 1 1 0.613 1 1 1 1
C11: Transnational governance 0 0 1 1 0 0 1 1 1 1
C12: Technology push 0 0 1 1 0 0 1 1 1 1
C13: Market-pull 0 0 1 1 0 0 1 1 1 1
C14: Carbon price and carbon market 0 0 1 1 0 0 1 1 1 1
C15: Technology transfer from developed to developing countries 0 0 1 1 0 0 1 1 1 1
C16: Regenerative and conscious capitalism 0 0 0 0 1 1.000 1 1.000 0 0
C17: Market transformation strategy 0 0 0 0 1 0.906 1 0.906 0 0
C18: Customer satisfaction 0 0 0 0 0 0.899 0 0.899 0 0
C19: Collaborative and constructive lobbying 0 0 0 0 0 0 0 0 0 0
C20: New conception of transparency 0 0 0 0 0 0.911 0 0.911 0 -0.903
C21: Greenwashing 0 0 0 0 0 -0.976 0 -0.976 1 0.916
C22: Economic value 0 0 0 0 0 0.967 0 0.967 0 0
C23: Environmental value 0 0 0 0 0 0.969 0 0.969 0 -0.909
C24: Social value 0 0 0 0 0 0.992 0 0.992 0 -0.905
C25: Sustainable production 0 0.974 0 0.976 0 0.976 0 0.976 0 0.976
C26: Sustainable consumption 0 0.886 0 0.886 0 0.966 0 0.966 0 0.886
Number of iterations 85 9 85 8 9

Note: IV=initial value, FV=final value

10

107
Providing financial incentives, subsidies, and installation rebates to end consumers can create markets
for the adoption of efficient technologies and practices (UNEP, 2018; Taylor, 2008). The high cost of
these technologies is a significant deterrent preventing their adoption. Penalties may also be used
strategically for greater deployment and diffusion of technologies (ibid).

3.3. Systemic Corporate Strategy: Another Enabler

Systemic corporate strategies such as regenerative and conscious capitalism, market transformation,
collaborative and constructive lobbying, a new conception of transparency, etc. are likely to enable
dematerialisation and decarbonisation measures (Fig. 2, Table 1: Scenario 3).

A responsible business strategy that imbibes the triple-bottom-line (environmental, social, and
economic goals) in unison is seen as a new market transformation strategy. Collaborative and
constructive lobbying also supports dematerialisation and decarbonisation measures (Fig. 1).
Collaborative and constructive lobbying is likely to benefit both competitors and the environment,
enabling shared knowledge and multi-level co-operation to permeate through supply chains and
horizontally across all sectors (Allwood et al., 2019). Besides, enabling corporate strategies would help
achieve high environmental and social values (Fig. 2).

3.4. Governance, Policies, and Regulations Along with Systemic Corporate Strategy: the
Greatest Enablers

FCM-based simulation results demonstrate that the combined measures of i) dematerialisation and
decarbonisation, ii) governance, policies and regulations (transnational governance, technology push,
market-pull, carbon price and carbon market, and technology transfer from developed to developing
countries); and iii) systemic corporate strategies (regenerative and conscious capitalism, and market
transformation strategy) will prove most useful for industrial system transition (Fig. 2, Table 1: Scenario
4).

Transitions to net-zero carbon emission are likely to be contested, non-linear, and require a multi-level
perspective policy approach that addresses a broad spectrum of social, political, cultural, and technical
changes as well associated phase-out policies and involve a wide range of actors, including civil society
groups, local authorities, labour unions, industry associations, etc. (Material Economics, 2019; Rogge
and Johnstone, 2017).

In a globalised world, the position of the state has weakened somewhat, as corporations play an
increasingly significant role in societal governance. Global corporations have played a major role in
identifying public policy standards for reducing GHG emissions. They have successfully shifted the
debate from a carbon tax to an emissions trading scheme with generous emission concessions that
allowed large corporations to make high profits while maintaining their business-as-usual approach
(Dorsey, 2007).

The industrial system transition to dematerialisation and decarbonisation, however, requires new forms
of multi-level and multi-stakeholder transnational governance arrangements. Andonova et al. (2009)
defined transnational governance as a ‘distinct form of global governance, consisting of transnational
actors operating in a political sphere in which public and private actors interact across national borders
and political jurisdictions.’ It authoritatively steers public goals (ibid). Transnational governance
involves a variety of state and non-state actors, each contributing different capacities and sources of
authority (ibid). There has been growing international collaboration between public and private
transnational actors, resulting in hybrid transnational governance networks (Table 2). Such networks
perform four distinct governance functions: incompatible, complementary, transitional, and
instrumental (Klijn and Skelcher, 2007). In the dematerialisation and decarbonisation space, we observe

11

108
hybrid (public-private) networks, engaged in public or private forms of transnational governance (Table
2).

Banerjee (2010) suggests the need for an enforceability mechanism for effective accountability of such
networks while Korbin (2009) proposes a transnational, multi-actor system of governance with a full
set of compliance-related norms and soft laws for ensuring accountability of actors in transnational
governance. We tried to analyse the accountability of leading transnational governance networks of
decarbonisation (climate change mitigation) space represented by monitoring mechanism, transparency
and representation of stakeholders based on the content analysis of disclosures on the official websites
of the respective partnerships (Table 2) as accountability is central to transnational legitimacy.

Most transnational governance networks are represented mainly by large players. The role of small
players like small and medium industries cannot be ruled out. Ostrom (2010) foresees polycentric
governance as more diverse, multi-leveled, with a greater emphasis on bottom-up initiatives that could
accommodate the interests of various actors.

3.5. Greenwashing: a Great Dragger

The FCM-based simulations also reveal that the deceptive corporate strategy (greenwashing) will not
only reduce the effectiveness of industrial system transition but also place the environmental and social
values in the negative zone (Fig. 2, Table 1: Scenario 5). A new conception of transparency may reduce
greenwashing (Wu et al., 2020) (Fig. 1). Increasing the transparency of environmental performance is
likely to play a crucial role in decreasing the incidence of greenwashing.

4. CONCLUSIONS

The industry as a whole is a very complex web of sectors, sub-sectors, inter-sectoral interactions, and
dependence. Interactions between dematerialisation, decarbonisation, systemic corporate strategies, and
governance and policies are far more complex. FCM-based simulations have proved handy in capturing
such complex interactions.

The findings of our FCM-based simulation help us decide that the technical measures of material
(including biomaterial) efficiency, circular economy, innovations in the energy sources (CDR
techniques, the role of hydrogen in the industry, etc.), and energy efficiency are necessary but not
sufficient conditions for industrial system transition leading into sustainable production and
consumption.

While materials’ efficiency and circular economy are incredibly crucial for dematerialisation, the role
of green innovations and green supply chain management cannot be overruled. Net-zero industrial
sector emissions require the provision of greenhouse gas emissions-free electricity, other energy
carriers, feedstock, and a combination of energy efficiency, reduced materials demand, improved
materials efficiency, a more circular economy, electrification, as well as carbon capture and use and
storage (CCUS).

109
Table 2: Global partnerships for decarbonisation and dematerialization: functions and accountability

Name of Type of Membership Main Function Accountability


Partnership partnership Monitoring Transparency Representation
Mechanism of stakeholders
REEEP Public-Private 359 including 45 national An open-ended initiative to facilitate multi-stakeholder Strong Medium High
(Hybrid) governments from all the cooperation in the areas of renewable energy,
continents; a cooperative decarbonisation, climate change, and sustainable
platform for more than 3500 development. It tries to reshape energy markets and develop
members the capacity for energy efficiency with partners
GMI Public-Private 45 partner countries and more Facilitates the exchange of information and knowledge to Strong High Medium
(Hybrid) than 700 project network advance methane mitigation in oil and gas, biogas,
members and coal mines sectors
Gold Standard Public-Private 16 institutional partners, 10 Facilitates decarbonisation goals through project-based Strong Low Medium
(Hybrid) strategic partners, 39 NGO interventions
supporters, and 42 network
platform partners
CDM Public-Private A large number of CDM To help developed countries fulfill their commitments to Very Strong High High
(Hybrid) projects with a network of reduce emissions and generate local, sustainable
institutions development solutions to developing countries
IPHE Governmental 20 partner countries Promotes hydrogen and fuel cell technologies across Moderate Medium to Low
Partnership applications and sectors to accelerate the transition to clean weak
and efficient energy system
CSLF Governmental 26 member governments (25 A collaborative effort to address the key technical, Moderate Medium to High
Partnership countries plus the European economic, and environmental obstacles in the development weak
Commission) and deployment of carbon capture and storage (CCS)
technologies
WBCSD Private-Private 200 leading private businesses Promotes sustainable businesses to accelerate the transition Strong High High
represented by CEOs to sustainable development
Renewable Energy and Energy Efficiency Partnership (REEEP); Global Methane Initiative (GMI); Clean Development Mechanism (CDM); International Partnership for Hydrogen and Fuel Cells
in the Economy (IPHE); Carbon Sequestration Leadership Forum (CSLF); World Business Council for Sustainable Development (WBCSD)

Source: Official websites of the respective partnerships


Note: Accountability dimensions are authors’ interpretation based on the content analysis of disclosures on the official websites of the respective partnerships

110
FCM-based simulation results demonstrate that the combined measures of (i) dematerialisation and
decarbonisation; (ii) governance, policies and regulations (transnational governance, technology push,
market-pull, carbon price and carbon market, and technology transfer from developed to developing
countries); and (iii) systemic corporate strategies (regenerative and conscious capitalism, and market
transformation strategy) are required for the stellar performance of industrial system transition to
dematerialisation and deep decarbonisation.

Policy approaches for dematerialisation and decarbonisation need to combine a mix of instruments
including carbon pricing, material efficiency and high-quality recycling policies, sectoral technology
roadmaps for new production processes, technology push and market-pull policies and planning. The
combination of these instruments depends on specific sectoral market barriers, technology maturity,
and local political and social acceptance.

The industrial system transition to dematerialisation and decarbonisation requires new forms of multi-
level and multi-stakeholder transnational governance arrangements, several of them are already
operating. These international collaborative network institutions have been increasingly engaged in
regulative, normative, cognitive, and identity-constructing governance, mostly in the decarbonisation
domain. However, these networks are mainly represented by large players. The participation of a large
number of small and medium industries and other small actors can be facilitated through polycentric
and trans-local governance arrangements.

REFERENCES
Abhishek, N., Divyashree, M.S., 2019. Environmental Friendly Supply Chain Management -A Perception Analysis. Journal of Supply Chain
Management Systems 8 (2): 45–51.
Åhman, M., Nilsson, L., Johansson, B., 2017. Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy
17 (5): 634–649. doi:10.1080/14693062.2016.1167009.
Allwood, J., Ashby, M., Gutowski, T., Worrell, E., 2011. Material efficiency: A white paper. Resources, Conservation and Recycling 55 (3):
362–381. doi:10.1016/j.resconrec.2010.11.002.
Allwood, J., Ashby, M., Gutowski, T., Worrell, E., 2013. Material efficiency: providing material services with less material production.
Philosophical Transactions of the Royal Society A 371.
Allwood, J., Dunant, C.F., Lupton, R., Cleaver, C., Serrenho, A., Azevedo, J., Horton, P., Clare, A., Low, H., Horrocks, I., Murray, J., Lin, J.,
Cullen, J., Ward, M., Salamati, M., Felin, T., Tim, I., Zhou, W., Hawkins, W., 2019. Absolute Zero. UK FIRES.
doi:10.17863/CAM.46075.
Andonova, L.B., Betsill, M.M., Bulkeley, H., 2009. Transnational climate governance. Global Environmental Politics 9: 52–73.
Banerjee, S.B., 2010. Governing the Global Corporation: A Critical Perspective. Business Ethics Quarterly 20 (2): 265–
274. doi:10.5840/beq201020219.
Bataille, C., Åhman, M., Neuhoff, K., Nilsson, L.J., Fischedick, M., Lechtenböhmer, S., Solano-Rodriquez, B., Denis-Ryan, A., Stiebert, S.,
Waisman, H., Sartor, O. Rahbar, S., 2018. A review of technology and policy deep decarbonization pathway options for making
energy intensive industry production consistent with the Paris Agreement. Journal of Cleaner Production 187: 960–973.
doi:10.1016/j.jclepro.2018.03.107.
Bleischwitz, R., Nechifor, V., Winning, M., Huang, B., Geng, Y., 2018. Extrapolation or saturation – Revisiting growth patterns, development
stages and decoupling. Global Environmental Change 48: 86–96. doi:10.1016/J.GLOENVCHA.2017.11.008.
Boyce, J., 2018. Carbon Pricing: Effectiveness and Equity. Ecological Economics 150: 52–61.
Carvalho, J.P., 2013. On the semantics and the use of fuzzy cognitive maps and dynamic cognitive maps in social sciences. Fuzzy Sets and
Systems 214: 6–19. doi:10.1016/j.fss.2011.12.009.
CDM, 2020. Clean Development Mechanism. https://cdm.unfccc.int [Accessed 25 July 2020].
CIWMB, 2003. Plastics White Paper—optimizing plastics use, recycling, and disposal in California. CA: White Paper. California Integrated
Waste Management Board (CIWMB). Publications Clearinghouse 916: 341–6306.
CSLF, 2020. Carbon Sequestration Leadership Forum. https://www.cslforum.org/cslf [Accessed 25 July 2020].
de Coninck, H., Revi, A., Babiker, M., Bertoldi, P., Buckeridge, M., Cartwright, A., Dong, W., Ford, J., Fuss, S., Hourcade, J.C., Ley, D.,
Mechler, R., Newman, P., Revokatova, A., Schultz, S., Steg, L., Sugiyama, T., 2018, Strengthening and implementing the global
response. In: Global warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial

111
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 [Masson- Delmotte, V., Zhai, P., Pörtner, H.O., Roberts, D., Skea,
J., Shukla, P.R., Pirani, A., Moufouma-Okia, W., Péan, C., Pidcock, R., Connors, S., Matthews, J.B.R., Chen, Y., Zhou, X., Gomis,
M.I., Lonnoy, E., Maycock, T., Tignor, M., Waterfield, T., (eds.)]. In Press.
De Jong, M.D.T., Harkink, K.M., Barth, S., 2017. Making Green Stuff? Effects of Corporate Greenwashing on Consumers. Journal of Business
and Technical Communication 32 (1): 77–112. doi:10.1177/1050651917729863.
Delmas, M., Burbano, V., 2011. The Drivers of Greenwashing. California Management Review. 54 (1): 64–87. doi:10.1525/cmr.2011.54.1.64.
Dorsey, M.K., 2007. Climate Knowledge and Power: Tales of Skeptic Tanks, Weather Gods, and Sagas for Climate
(In)justice. Capitalism, Nature, Socialism 18 (2): 7–21.
Fischedick, M., Roy, J., Abdel-Aziz, A., Acquaye, A., Allwood, J.M., Ceron, J.-P., Geng, Y., Kheshgi, H., Lanza, A., Perczyk, D., Price, L.,
Santalla, E., Sheinbaum, C., Tanaka, K., 2014. Industry. In: Climate Change 2014: Mitigation of Climate Change. Contribution of
Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., Pichs-
Madruga, R., Sokona,Y., Farahani, E., Kadner, S., Seyboth, K., Adler, A., Baum, I., Brunner, S., Eickemeier, P., Kriemann, B.,
Savolainen, J., Schlömer, S., von Stechow, C., Zwickel, T., Minx, J.C., (eds.)]. Cambridge University Press, Cambridge, United
Kingdom and New York, NY, USA.
Fullerton, J., 2015. Regenerative capitalism - How universal principles and patterns will shape our new economy. Capital Institute.
Gardner, T.A. Benzie, M., Börner, J., Dawkins, E., Fick, S., Garrett, R., Godar, J., Grimard, A., Lake, S., Larsen, R.K., Mardas, N., McDermott,
C.L., Meyfroidt, P., Osbeck, M., Persson, 2019. Transparency and sustainability in global commodity supply chains. World
Development 121: 163–177. doi:10.1016/j.worlddev.2018.05.025.
Garmulewicz, A., Holweg, M., Veldhius, H. Yang, A., 2018. Disruptive Technology as an Enabler of the Circular Economy: What Potential
Does 3D Printing Hold?. California Managemnet Review 60 (3): 112–132. doi:10.1177/0008125617752695.
Ghisetti, C., 2017. Demand-pull and environmental innovations: Estimating the effects of innovative public procurement. Technological
Forecasting & Social Change 125: 178–187. doi:10.1016/j.techfore.2017.07.020.
GMI, 2020. Global Methane Initiative. https://www.globalmethane.org [Accessed 25 July 2020].
Gold Standard, 2020. https://www.goldstandard.org [Accessed 25 July 2020].
Goodman, J., Korsunova, A., Halme, M., 2017. Our collaborative future: activities and roles of stakeholders in sustainability-oriented
innovation. Business Strategy and the Environment 26 (6). 731–753. doi:10.1002/bse.1941.
Grubb, M., McDowall, W., Drummond, P., 2017. On order and complexity in innovations systems: Conceptual frameworks for policy mixes
in sustainability transitions. Energy Research & Social Science 33: 21–34. doi:10.1016/J.ERSS.2017.09.016.
Grubler, A., Wilson, C., Bento, N., Boza-Kiss, B., Krey, V., McCollum, D., Rao. N., Riahi, K., Rogelj, J., De Stercke, S., Cullen, J., Frank,
S., Fricko, O., Guo, F., Gidden, M., Havlík, P., Huppmann, D., Kiesewetter, G., Rafaj, P., Schoepp, W., Valin, H., 2018. A low
energy demand scenario for meeting the 1.5°C target and sustainable development goals without negative emission technologies.
Nature Energy 3: 515–527. doi:10.1038/s41560-018-0172-6.
Hansen, E., Lüdeke-Freund, F., Quan, X., West, J., 2019. Cross-National Complementarity of Technology Push, Demand Pull, and
Manufacturing Push Policies: The Case of Photovoltaics. Institute of Electrical and Electronics Engineers 66 (3): 381–397.
doi:10.1109/TEM.2018.2833878
Heshmati, A., 2018. A review of the circular economy and its implementation. International Journal of Green Economics 11 (3/4). 251.
doi:10.1504/IJGE.2017.089856.
Hildingsson, R., Kronsell, A., Khan, J., 2019. The green state and industrial decarbonisation. Environmental Politics 28 (5): 909–928.
doi:10.1080/09644016.2018.1488484.
Hoffman, A., 2018. The Next Phase of Business Sustainability. Stanford Social Innovation Review 16 (2): 34–39.
Holland, T., Smit, B., 2014. Recent climate change in the Prince Edward County winegrowing region, Ontario, Canada: implications for
adaptation in a fledgling wine industry. Regional Environmental Change 14: 1109–1121. doi:10.1007/s10113-013-0555-y.
Hosseini, S.E., 2020. An outlook on the global development of renewable and sustainable energy at the time of COVID-19. Energy Research
& Social Science 68: 101633. doi:10.1016/j.erss.2020.101633.
Hurlimann, A.C., Browne, G.R., Warren-Myers, Francis, V., 2018. Barriers to climate change adaptation in the Australian construction
industry – Impetus for regulatory reform. Building and Environment 137: 235–245. doi:10.1108/JPIF-12-2019-0161.
IEA, 2017. Energy Technology Perspectives 2017. https://webstore.iea.org/download/direct/1058. [Accessed 11 July 2020].
IEA, 2019. Material efficiency in clean energy transitions. International Energy Agency. OECD. DOI:10.1787/aeaaccd8-en.
IEA, 2020. Global Energy Review 2020: The impacts of the Covid-19 crisis on global energy demand and CO2 emissions. Paris: International
Energy Agency.
IPHE, 2020. International Partnership for Hydrogen and Fuel Cells in the Economy. https://www.iphe.net [Accessed 25 July 2020].
Jacobsson, S., Bergek, A., 2011. Innovation system analyses and sustainability transitions: Contributions and suggestions for research.
Environmental Innovation and Societal Transitions 1: 47–51. doi:10.1016/j.eist.2011.04.006.
Kang, D., Lee, D., 2016. Energy and environment efficiency of industry and its productivity. Journal of Cleaner Production 135 (1): 184–
193. doi:10.1016/j.jclepro.2016.06.042.
Kim, E., Lyon, T., 2015. Greenwash vs. Brownwash: exaggeration and undue modesty in corporate sustainability disclosure. Organization
Science 26 (3): 705–723. doi:10.1287/orsc.2014.0949.
Klijn, E.-H., Skelcher, C., 2007. Democracy and Governance Networks: Compatible or Not? Public Administration 85 (3): 587–608.
doi:10.1111/j.1467-9299.2007.00662.x.

112
Köhler, J., Geels, F. W., Kern, F., Markard, J., Wieczorek, A., Alkemade, F., Avelino, F., Bergek, A., Boons, F., Fünfschilling, L., Hess, D.,
Holtz, G., Hyysalo, S., Jenkins, K., Kivimaa, P., Martiskainen, M., McMeekin, A., Mühlemeier, M. S., Nykvist, B., Onsongo, E. Pel,
B., Raven, R., Rohracher, H., Sandén, B., Schot, J., Sovacool, B., Turnheim, B., Welch, D., Wells, P., 2019. An agenda for
sustainability transitions research: State of the art and future directions. Environmental Innovation and Societal Transitions 31: 1–32.
doi:10.1016/j.eist.2019.01.004.
Korbin, S.J., 2009. Private Political Authority and Public Responsibility: Transnational Politics, Transnational Firms, and Human
Rights. Business Ethics Quarterly 19(3): 349–74.
Kosko, B., 1986. Fuzzy Cognitive Maps. International Journal of Man-Machine Studies 24: 65–75.
Kosko, B., 1987. Adaptive bidirectional associative memories. Applied optics 26 (23): 4947–4960.
Lema, R., Lema, A., 2012. Technology transfer? The rise of China and India in green technology sectors. Innovation and Development 2 (1):
23–44. doi:10.1080/2157930X.2012.667206.
Lereboullet, A., Beltrando, G., Bardsley, D., 2013. Socio-ecological adaptation to climate change: A comparative case study from the
Mediterranean wine industry in France and Australia. Agriculture, Ecosystems & Environment 164: 273–285.
doi:10.1016/j.agee.2012.10.008.
Linnenluecke, M.K., Griffiths, A., Winn, M.I., 2013. Firm and industry adaptation to climate change: a review of climate adaptation studies
in the business and management field. WIREs Climate Change. doi:10.1002/wcc.214.
Long, T.B., Looijen, A., Blok, V., 2018. Critical success factors for the transition to business models for sustainability in the food and beverage
industry in the Netherlands. Journal of Cleaner Production 175: 92–95. doi:10.1016/j.jclepro.2017.11.067.
Mackey, J., Sisodia, R., 2014. Conscious Capitalism. Harvard Business Review Press.
Marchi, V., 2012. Environmental innovation and R&D cooperation: Empirical evidence from Spanish manufacturing firms. Research Policy
41: 614–623. doi:10.1016/j.respol.2011.10.002.
Marquis, C., Toffel, M., Zhou, Y., 2016. Scrutiny, norms, and selective disclosure: a global study of greenwashing. Organization Science 27
(2): 483–504. doi:10.1287/orsc.2015.1039.
Material Economics, 2019. Industrial Transformation 2050: Pathways to net-zero emissions from EU Heavy Industry.
https://materialeconomics.com/publications/industrial-transformation-2050. [Accessed 11 July 2020].
Morone, P., Falcone, P.M., Lopolito, A., 2019. How to promote a new and sustainable food consumption model: a fuzzy cognitive map study.
Journal of Cleaner Production 208: 563–574. doi:10.1016/j.jclepro.2018.10.075.
Mosquera, M., 2019. Banning Plastic Straws: The Beginning of the War against Plastics. Earth Jurisprudence and Environmental Justice
Journal 9: 5–31.
Nápoles, G., Papageorgiou, E., Bello, R., Vanhoof, K., 2016. On the convergence of sigmoid Fuzzy Cognitive Maps”. Information Sciences
349: 154–171. doi:10.1016/j.ins.2016.02.040.
Newell, S., Goldsmith, R., Banzhaf, E., 2015. The Effect of Misleading Environmental Claims on Consumer Perceptions of Advertisements.
Journal of Marketing Theory and Practice 48–60.
OECD, 2019. Moving to sustainable industrial production. Accelerating Climate Action: Refocusing Policies through a Well-being Lens, 65–
88.
Olatunji, O.O., Ayo, O.O., Akinlabi, S., Ishola, F., Madushele, N., Adedeji, P.A., 2019. Competitive advantage of carbon efficient supply
chain in manufacturing industry. Journal of Cleaner Production 238 (20): 117937. doi:10.1016/j.jclepro.2019.117937.
Ostrom, E., 2010. Polycentric systems for coping with collective action and global environmental change. Global Environmental Change 20:
550–557. doi:10.1016/j.gloenvcha.2010.07.004.
Özesmi, U., Özesmi, L.S., 2004. Ecological models based on people’s knowledge: a multi-step fuzzy cognitive mapping approach. Ecological
Modelling 176 (1-2): 43–64. doi:10.1016/j.ecolmodel.2003.10.027.
Petrides, D., Papacharalampopoulos, A., Stavropoulos, P., Chryssolouris, G., 2018. Dematerialisation of products and manufacturing-
generated knowledge content: relationship through paradigms. International Journal of Production Research 56 (1-2): 86–96.
doi:10.1080/00207543.2017.1401246.
Potting, J., Hanemaaijer, A., Delahaye, R., Hoekstra, R., Ganzevles, J., Lijzen, J., (eds.), 2018. Circular Economy: What we want to know and
can measure: Framework and baseline assessment for monitoring the progress of the circular economy in the Netherlands. The Hague:
PBL Netherlands Environmental Assessment Agency.
Randolph, J., Masters, G.M., 2018. Market Transformation to Sustainable Energy. In: Energy for Sustainability. Island Press. Washington,
DC. doi:10.5822/978-1-61091-821-3_16.
Rankin, W., 2014. Sustainability. In: Treatise on Process Metallurgy Volume 3: Industrial Processes [Seetharaman, S. (ed.)]. Elsevier. 1376–
1424.
REEEP, 2020. Renewable Energy and Energy Efficiency Partnership. https://www.reeep.org [Accessed 25 July 2020].
Ritchie, H., Roser, M., 2019. CO₂ and Greenhouse Gas Emissions. https://ourworldindata.org/co2-and-other-greenhouse-gas-emissions
[Accessed 28 April 2020].
Rizos, V., Behrens, A., Van der Gaast, W., Hofman, E., Ioannou, A., Kafyeke, T., Flamos, A., Rinaldi, R., Papadelis, S., Hirschnitz-Garbers,
M., Topi, C., 2016. Implementation of Circular Economy Business Models by Small and Medium-Sized Enterprises (SMEs): Barriers
and Enablers. Sustainability 8 (11): 1212. doi:10.3390/su8111212.
Rockström, J., Gaffney, O., Rogelj, J., Meinshausen, M., Nakicenovic, N., Schellnhuber, H.J., 2017. A roadmap for rapid decarbonization -
Emissions inevitably approach zero with a carbon law. Science 355 (6331): 1269–1271. doi:10.1126/science.aah3443.
Rogge, K.S., Johnstone, P., 2017. Exploring the role of phase-out policies for low-carbon energy transitions: The case of the German
Energiewende. Energy Research & Social Science 33: 128–137. doi:10.1016/j.erss.2017.10.004.

113
Ryan, L., Moarif, S., Levina, E., Baron, R., 2011. Energy efficiency policy and carbon pricing. Paris: International Energy Agency.
Schmalensee, R., Stavins, R., 2017. Lessons Learned from Three Decades of Experience with Cap and Trade. Review of Environmental
Economics and Policy 11: 59–79. doi:10.1093/reep/rew017.
Schwarz, M., Nakhle, C., Knoeri, C., 2020. Innovative designs of building energy codes for building decarbonization and their implementation
challenges. Journal of Cleaner Production 248 (1): 119260. doi:10.1016/j.jclepro.2019.119260.
Singh, P.K., Chudasama, H., 2017a. Pathways for drought resilient livelihoods based on people’s perception. Climatic Change 140: 179–193.
doi:10.1007/s10584-016-1817-8.
Singh, P.K., Chudasama, H., 2017b. Assessing impacts and community preparedness to cyclones: a fuzzy cognitive mapping approach.
Climatic Change 143: 337–354. doi:10.1007/s10584-017-2007-z.
Singh, P.K., Chudasama, H., 2020. Evaluating Poverty Alleviation Strategies in a Developing Country. PLoS ONE 15 (1): e0227176.
doi:10.1371/journal.pone.0227176.
Singh, P.K., Papageorgiou, K., Chudasama, H., Papageorgiou, E.I., 2019. Evaluating the Effectiveness of Climate Change Adaptations in the
World’s Largest Mangrove Ecosystem. Sustainability 11 (23): 6655. doi:10.3390/su11236655.
Söderholm, P., Tilton, J.E., 2012. Material efficiency: An economic perspective. Resources, Conservation and Recycling 61: 75–82.
doi:10.1016/j.resconrec.2012.01.003.
Solana-Gutiérrez, J., Rincón, G., Alonso, C., García-de-Jalón, D., 2017. Using fuzzy cognitive maps for predicting river management
responses: A case study of the Esla River basin, Spain. Ecological Modelling 360: 260–269. doi:10.1016/j.ecolmodel.2017.07.010.
Stylios, C.D., Georgopoulos, V.C., Groumpos, P.P., 1997. The use of fuzzy cognitive maps in modeling systems. In: Proceedings of 5th IEEE
Mediterranean conference on control and systems, Paphos, Cyprus, 21–23.
Sun, H., Edziah, B., Sun, C., Kporsu, A., 2019. Institutional quality, green innovation and energy efficiency. Energy Policy 135.
doi:10.1016/j.enpol.2019.111002.
Tanaka, K., 2011. Review of policies and measures for energy efficiency in industry sector. Energy Policy 39: 6532–6550.
Taylor, M., 2008. Beyond technology-push and demand-pull: Lessons from California's solar policy. Energy Economics 30: 2829–2854.
UNEP, 2018. Single-Use Plastics: A Roadmap for Sustainability. United Nations Environment Programme. (Rev. ed., pp. vi; 6).
UNIDO, 2013. Lima Declaration: Towards inclusive and sustainable industrial development. In: Fifteenth Session of the UNIDO General
Conference. Lima, 2013. UNIDO.
Waddock, S., 2020. Achieving sustainability requires systemic business transformation. Global Sustainability 3: 1–12. doi:10.1002/bse.1970.
WBCSD, 2020. World Business Council for Sustainable Development. https://www.wbcsd.org [Accessed 25 July 2020].
Wei, W., Chen, D., Hu, D., 2016. Study on the Evolvement of Technology Development and Energy Efficiency—A Case Study of the Past
30 Years of Development in Shanghai. Sustainability 8 (5). 457. doi:10.3390/su8050457.
Worrell, E., Allwood, J., Gutowski, T., 2016. The Role of Material Efficiency in Environmental Stewardship. Annual Review of Environment
and Resources 41 (7): 1–24. doi:10.1146/annurev-environ-110615-085737.
Wu, Y., Zhang, K., Xie, J., 2020. Bad Greenwashing, Good Greenwashing: Corporate Social Responsibility and Information Transparency.
Management Science. doi:10.1287/mnsc.2019.3340.
Xiang, W., Ming, C., 2011. Implementing extended producer responsibility: vehicle remanufacturing in China. Journal of Cleaner Production
19: 680–686. doi:10.1016/j.jclepro.2010.11.016.
Ziv, G., Watson, E., Young, D., Howard, D.C., Larcom, S.T., Tanentzap, A.J., 2018. The potential impact of Brexit on the energy, water and
food nexus in the UK: A fuzzy cognitive mapping approach. Applied Energy 210: 487–498. doi:10.1016/j.apenergy.2017.08.033.

114
10. Pathways for Climate Resilient Development: Human Well-Being within a Safe and
Just Space in the 21st Century1

Pramod K. Singh and Harpalsinh Chudasama

Abstract

The concept of climate resilient development pathways introduced in IPCC’s Fifth Assessment Report
remains poorly conceptualised. We have attempted to deepen the conceptualisation of climate resilient
development (CRD) or climate compatible development, while charting its pathways through fuzzy
cognitive maps (FCM)-based simulations aided by knowledge based on stakeholders’ insights. We
conceptualise CRD as a development embracing mitigation, adaptation and inclusive sustainable
development to advance planetary health and well-being for all. The FCM-based simulations
demonstrate that appropriate enabling conditions are critical to the achievement of CRD, the most
important of them being (i) the ethics, values, and worldviews shaping CRD’s directions by framing
appropriate climate narratives and action; (ii) partnerships and commitment to finance and technology
by the governments; (iii) interactions between the actors and arenas of engagement facilitating CRD
decisions and actions; and (iv) dimensions of governance at multiple levels involving policy,
institutions and practice. Citizens’ defence against climate change as a civil right, along with
planetary health and well-being, demands synergies while implementing mitigation, adaptation and
sustainable development. Short-term decisions and actions related to mitigation, adaptation, and
sustainable development could have long-term effects on CRD pathways. CRD could entail a deep
societal transformation to ensure universal well-being. The findings of this research could have
profound implications for multilateral negotiations.

Keywords: Climate resilient development, Climate compatible development, governance, planetary health, sustainable
development, well-being

1. Introduction

IPCC’s Fifth Assessment Report (AR5) introduced the concept of climate-resilient development
pathways (CRDPs) as ‘development trajectories that combine adaptation and mitigation to realise the
goal of sustainable development’ (Denton et al., 2014, p.1104). The emerging climate resilient
development literature highlights the urgency of climate actions facilitating substantial reductions in
greenhouse gas emissions while pushing gains in human and natural system resilience (Haines et al.,
2017; Shindell et al., 2017; Nerini et al., 2019). Mitchell and Maxwell (2010) coined the term climate
compatible development (CCD) to describe a ‘triple win’ scenario showcasing adaptation, mitigation,
and development actions. Interdependence between development trends, climate risk, and mitigation
and adaptation action are central to the definition of CRDPs (Fankhauser and McDermott, 2016).

In addition, the United Nations 2030 Agenda aims at time-bound targets for prosperity, people, planet,
peace and partnership with the commitment that no one will be left behind in achieving the
Sustainable Development Goals (SDGs) (UN, 2015) set out by the former. The Paris Climate
Agreement provided a climate resilient development paradigm by establishing a mitigation-centered
target of ‘limiting warming well below 2°C above pre-industrial level’ to reduce risks associated with
higher warming (Craft and Fisher, 2018). Different climate scenarios (1.5°C, 2°C, or even higher

1
Under review with Global Environmental Change, reproduced for class discussion only.

115
ones) could potentially create various degrees of impacts across sectors and regions (Solecki et al.,
2018). The current nationally determined contributions of mitigation pledge would have a slim chance
of reaching even the 2°C target despite the great hopes on carbon dioxide removal efforts (Grassi et
al., 2017). Mitigation risks have jeopardised national and local efforts to eradicate poverty, end
hunger, reduce inequalities, and achieve several SDGs, especially in the Global South (Seneviratne et
al., 2018). Thus, the underlying conceptualisation of CRD appears inadequate.

Meeting SDGs will not only enhance ecological integrity and human well-being but also help improve
mitigation and adaptation planning and execution (Conway et al., 2015; Griscom et al., 2017; Allen et
al., 2018; Roy et al., 2018). However, pursuing SDG priorities to promote ecological security and
human well-being highlights the complex dynamics of synergies and trade-offs between socio-
economic development and climate change responses (Santika et al., 2019). The integration of climate
justice (Patterson et al., 2018), ethics, values (O’Neill et al., 2015; O’Brien, 2018), and worldviews
(Hayward, 2012; Wolf et al., 2013) plays a crucial role in shaping the ethical goal of climate resilient
development.

Extant literature indicates that multiple players and a plurality of knowledge are required to inform
inequitable partnerships and improve structures and processes for climate resilience (Klenk et al.,
2017; Hulme, 2018). Climate resilience is feasible only if the SDGs have been fulfilled (de Loma-
Osorio, 2016). There are, however, inherent contradictions in fulfilling all SDGs (Spaiser et al., 2017).
The United Nations Economic and Social Commission reports that since progress has either stagnated
or shifted in a reverse direction for over 50% SDGs, Asia and Africa will not be able to achieve them
(UN ESCAP, 2019).

While deepening the conceptualisation of climate resilient development (CRD) and climate resilient
development pathways (CRDPs), this paper suggests pathways for CRD with the aid of fuzzy
cognitive maps (FCM)-based simulations that are used to model the interactions occurring within a
complex and hard-to-model socio-ecological system. The paper attempts to explore the following two
major research questions.

• What are the relative roles of adaptation, mitigation, and SDGs in attaining CRD?
• How do various enabling conditions help in charting CRD?

1.2. Rationale for FCM approach

Quantitative models like integrated assessment models, while acknowledging real-world dynamics
(Geels et al., 2016), are not designed to reflect the multiple dimensions of well-being, poverty, and
inequalities of various climate policies and pathways (O’Neill et al., 2018). In this context, FCM,
which relies on knowledge gained from experts on the basis of their experience, knowledge, and
perception, comes in handy during the quantitative modelling of interactions occurring within
complex and hard-to-model qualitative socio-ecological systems (Özesmi and Özesmi, 2004; Singh
and Chudasama, 2017; 2020; 2021; Gray et al., 2019; Fonseca et al., 2020; Pereira et al., 2020).

2. Theoretical Foundation: Deepening the Conceptualisations of CRD and CRDPs

The current conceptualisation of CRD and CRDP evolved through the lead authors’ meetings of
chapter 18 of the IPCC Working Group II of its Sixth Assessment Report. The aim of adaptation,
mitigation, and sustainable development is to promote human well-being on a healthy planet.

116
2.1. Human well-being

Human well-being is an ‘abstraction used to refer to a description of the state of individuals’ life
situations (Mcgillivray and Clarke, 2006) and represents an optimal experience and functioning. Two
schools of human well-being—hedonic and eudaimonic—have been locked in a debate in this
context. Hedonic well-being reflects a subjective condition of human motivation and is linked to one’s
assumed degrees of happiness or pleasure. It has gained traction in the psychological evaluation of
‘subjective well-being’ including happiness and life satisfaction, assuming that the individual is
driven to enhance personal freedom and self-preservation (Lamb and Steinberger, 2017; Ganglmair-
Wooliscroft and Wooliscroft, 2019). Eudaimonic well-being focusses on the individual in a broader
context, associating happiness with virtue (Sirgy, 2012), encouraging social institutions and political
systems to thrive while enabling individuals to flourish. Thus, it fixates not so much on the outcome
but on the process of living well. Based on Self-Determination Theory, Ryan and Deci (2008)
propounded four motivational concepts for eudaimonic living: (i) pursuing intrinsic goals and values;
(ii) behaving in autonomous, or volitional ways; (iii) acting with a sense of full awareness; and (iv)
exhibiting behavior to fulfill basic psychological needs of relatedness, and autonomy. Eudaimonic
well-being embraces development approaches including capabilities (Sen, 1985) and human needs
(Doyal and Gough, 1991). Such needs subsume survival (physiology, safety, security), the
psychological (love, belonging, esteem, self-actualisation), and spiritual (self-transcendence)
(Dominati et al., 2010).

Human well-being can best be seen as a multidimensional phenomenon that involves elements of both
hedonic and eudaimonic concepts of well-being. It not only fulfills a range of human needs but also
strengthens the ability to endevour towards a satisfying life. Well-being components involve health,
happiness, meaningful work and social relationships, freedom and liberties, and determinants like
food, shelter, water, and access to knowledge and information (Dasgupta, 2001). They concentrate on
intrinsic goals and values, relatedness, connectedness with nature, and the development of human
potential. Sen (1993) sees human capabilities relevant for human well-being as physical life, health,
physical integrity, senses, imagination, thinking, emotions, reflection, and relationship with other
species along with political and material control over one’s own environment. SDGs 1 (end poverty),
2 (end hunger), 3 (good health), 4 (equitable quality education), 5 (gender equality and women
empowerment), 10 (reduced inequality) and 16 (peace, justice, and strong institutions) are the main
facets of enhancing human well-being.

The Millennium Ecosystem Assessment (2005) proposed five dimensions to measure human well-
being: (i) basic material for a good life (food, water, shelter, adequate income, household assets), (ii)
health, (iii) good social relations (develop institutional linkages that create social capital, show mutual
respect, have good gender and family relations), (iv) security, and (v) freedom of choice. The Stiglitz-
Sen-Fitoussi Commission set-up by the French government suggested eight key dimensions of human
well-being: (i) material living standards (income, consumption, and wealth), (ii) health, (iii)
education, (iv) freedom of personal activities including work, (v) political voice and governance, (vi)
social connections and relationships, (vii) present and future social and physical environmental
conditions, and (viii) physical and economic securities (Stiglitz et al., 2009). The resilience of the
community will be extremely crucial for well-being in the context of climate change. Thus, human
well-being signifies equity, justice, connectedness to nature and society, resilience, fulfillment,
esteem, ultimate pleasure, self-actualisation, and self-transcendence.

117
2.2. Planetary health

Planetary health is a growing concept focused on the interdependence of human, animal, and
environmental health and well-being. It is defined as the ‘health of human civilisation and the state of
the natural systems on which it depends’ (Horton et al., 2014). Rockstrom et al. (2009) identified nine
‘planetary boundaries’ representing ‘thresholds below which humanity can safely operate and beyond
which the stability of planetary-scale systems cannot be relied upon.’ Anthropogenic perturbation
levels of four of the Earth system processes/ features (climate change, biosphere integrity,
biogeochemical flow, and land system change) have been identified to exceed the ‘planetary
boundaries’ (Steffen et al., 2015). Planetary health signifies restricting climate change and,
maintaining regenerative ecosystem processes to ensure that the ecosystem delivers its services
adequately. Both planetary health and human well-being are profoundly ethical issues as they relate to
how humans can live a healthy life without endangering ecological security.

2.3. Designing the assessment framework

The SDG and Paris Climate Agreement goals are interdependent (Nilsson et al., 2016) with dynamic
interconnections between human, technological and natural processes (Sachs et al., 2019). The
climate-human-environment system is extremely complex and dynamic (Singh and Nair, 2014), with
the occurrence of complex cross-scale social-ecological-economic interactions (Sellberg et al., 2018).
Hence, we chose to model the complete climate-human-environment system to understand the
relationships between adaptation, mitigation and SDGs, which ultimately influence planetary health
and human well-being. These include: (i) key drivers influencing climate processes; (ii) climate
processes; (iii) climate change; (iv) adverse impacts; (v) losses and damages; (vi) key risks; (vii)
shocks; (viii) adaptation; (ix) mitigation; (x) sustainable development; (xi) enabling conditions; and
(xii) planetary health and human well-being (Fig. 1).
Preceding sections have already addressed issues related to planetary health and human well-being.
Besides, the shocks and enabling conditions in our assessment framework are somewhat external to
the climate-development system, awaiting further explanation. Here shocks are significant
occurrences that could potentially derail the development trajectory. We have experienced COVID-
19, a recent pandemic that triggered unparalleled human, economic, and livelihood devastation across
the globe. Synthetic biology holds the ability to generate bacteria and viruses with novel and deadly
characteristics. Solar radiation management has the potential to cause regional ecological disruptions.
In the arsenals of nine possessor countries (Kristensen and Norris, 2020), even a small proportion of
the projected 13,410 nuclear arms could theoretically end civilisation within one day.
We would like to provide definitional clarifications of some of the less understood enabling
conditions, namely: (i) ethics, values, and worldviews; (ii) governance; (iii) actors and arenas of
engagement; and (iv) societal transformation.
While ethics represent noble visions and high ideals (Hayward, 2012), values reflect intrinsically
desirable principles or qualities. Values ‘serve as standards or criteria to guide not only action but also
judgment, choice, attitude, evaluation, argument, exhortation, rationalisation, and, one might add,
attribution of causality.’ (Rokeach, 1979). Schwartz (1994) identified 10 types of basic and universal
values—security, tradition, conformity, power, achievement, hedonism, stimulation, self-direction,
universalism, and benevolence—and the characteristic motivations that organise them into value
systems. Worldview is a synthesis of the value orientation of an individual—‘the lens through which
the world is seen’ (van Egmond and de Vries, 2011).

118
Fig. 1: Assessment framework for climate resilient development

119
Governance, which determines how collective interests are defined, negotiated and pursued, and how
economy and society are steered or governed (Sørensen and Torfing, 2018), is one of the most
important enabling conditions in the development process. Governance actors (government, industry,
media, civil society, and science) make social choices either independently or collectively (Sørensen
and Torfing, 2018) across political, economic, socio-cultural, ecological, community, and knowledge
and technology arena of engagements.

IPCC (2019) defines societal transformation as ‘a profound and often deliberate shift initiated by
communities toward sustainability, facilitated by changes in individual and collective values and
behaviours, and a fairer balance of political, cultural, and institutional power in society.’

2.4. Working definition of CRD and CRDP

While explaining sustainable development, Daly (1995) put it elegantly: ‘All important concepts are
dialectically vague at the margins’ and the same can be said for human well-being. Human well-being
not only involves equity, justice, connectedness to nature and society but also embraces Maslow’s
higher-order needs of fulfillment, esteem, ultimate pleasure, self-actualisation, and self-transcendence.
Besides seeing SDGs as short-term (2030) goals for prosperity, people, planet, peace and partnership,
we considered human well-being on a healthy planet as CRD’s ultimate objective.
Accordingly, we propose climate resilient development (CRD) or climate compatible development
(CCD) as a development embracing mitigation, adaptation, and inclusive sustainable development to
advance planetary health and well-being for all. Accordingly, climate-resilient development
pathways (CRDPs) are alternate trajectories of climate resilient development.

3. Methods

Fuzzy cognitive mapping is a participatory modelling technique in which participants create a ‘fuzzy
cognitive map’ reflecting their perceived cause-and-effect relationships in a complex system (Fonseca
et al., 2020; Pereira et al., 2020; Singh and Chudasama, 2020; 2021). The approach helps visualise
how interrelated variables/ factors/ concepts (referred to as ‘concepts’ in this study) affect one another
and represent feedback within the system (Nápoles et al., 2016; Singh and Chudasama, 2017; 2020;
2021; Gray et al., 2019; Fonseca et al., 2020). Here, participants focus on the cause-and-effect
relationships between the qualitative concepts and generate quantitative data dependent on their
experiences, knowledge, and perceptions (Özesmi and Özesmi, 2004; Reckien, 2014; Singh and Nair,
2014; Singh and Chudasama, 2017; 2020; 2021; Fonseca et al., 2020; Pereira et al., 2020). The
following is a multi-step process of obtaining individual FCMs from the study participants and
analysing them.

3.1. Knowledge elicitation

FCM-based literature describes various approaches to designing participatory FCMs, including


‘through literature review’, the ‘Delphi method’, ‘survey’, or ‘facilitating participants directly’ to
construct an FCM (Gray et al., 2019), which can be classified under two categories: ‘pre-designed
concept’ and ‘open-concept design’ (Singh and Chudasama, 2020). We adopted a ‘pre-designed
concept’ approach through a literature review (Morone et al., 2019). Discussions during the IPCC lead
authors’ meeting helped to sharpen the concept maps.

In order to model the complex social-ecological-economic dynamics occurring within the complex
climate-human-environment system centred on our conceptualisation of climate resilient
development, we identified 12 major dimensions (concepts): key drivers and development trends,

120
climate processes, climate change, adverse impacts, losses and damages, key risks, shocks, adaptation,
mitigation, sustainable development, enabling conditions, and planetary health and human well-being
(Fig. 1). The literature review (supplementary material 1) and informal discussions with IPCC authors
helped us identify 107 sub-concepts within these 12 concepts (supplementary material 2). The
nomenclature of several sub-concepts (detailed concepts) evolved over time. Initially, we began with
97 sub-concepts, which went up to 107.

In order to gather the views of researchers, academicians, and IPCC authors of the ongoing Sixth
Assessment on climate-resilient development pathways we designed three instruments/ research
protocols depicting (a) the causal relationships of 12 main concepts, (b) the causal relationships of
107 sub-concepts, and (c) prioritisation exercise of 107 sub-concepts based on their relative
importance in the system. The survey instruments/ research protocol ‘a’ and ‘c’ were merged into a
graphic (Supplementary material 2). Besides, all the three instruments had tabular versions: causal
relationship statement tables for instruments ‘a’ and ‘b’ and prioritisation table for instrument ‘c’. We
have provided the relationship statement table for sub-concepts (instrument ‘b’) in supplementary
material 3.

For several days at the IPCC meetings' side-lines, particularly in the evenings, we held informal
discussions, initially, with over 70 authors in small groups of 10-17 participants. Besides sharing the
instruments via emails, we circulated the printed instruments/ research protocol ‘a’ and ‘c’ merged in
a graphic (supplementary material 2) instrument ‘b’ printed table (supplementary material 3) during
the informal meetings. During these interactions, we received suggestions regarding the inclusion of
sub-concepts. Many IPCC authors completed instruments ‘a’ and ‘c’ on the printed graphic and
instrument ‘b’ on the printed table, returning them mostly the next day or after a few days, while
others sent scanned copies later.

In addition, we distributed the survey instruments and definitions of the terms to the domain experts,
including researchers, academicians, and IPCC authors for their views on climate-resilient
development pathways. We also requested them to incorporate additional relationship/s in the
instruments relying on their experience, knowledge and perceptions. Several participants
communicated that the dimensions were very extensive, with no addition required.

During the information elicitation phase, each participant was asked to provide relationship weights of
the main concepts on a scale of 1–10, depicting the strength of the relationship between the concepts
regarding their importance to climate-resilient development, with 1 being the weakest and 10 the
strongest (Özesmi and Özesmi, 2004; Singh and Chudasama, 2017; Ziv et al., 2018; Pereira et al.,
2020). Positive or negative interaction between two concepts Ci and Cj indicate the nature of the
relationship weights, i.e., an increase (or decrease) in concept Ci will result in an increase (or
decrease) in concept Cj (Özesmi and Özesmi, 2004; Singh and Chudasama, 2017; Ziv et al., 2018;
Morone et al., 2019; Pereira et al., 2020). The participants were asked to assign a zero value when
they did not agree with a relationship (Gray et al., 2019). We did not obtain a zero value, however.
The participants were communicated that the values ranging 1–2 represented very low, 3–4 low, 5–6
moderate, 7–8 high, and 9–10 very high strengths (Özesmi and Özesmi, 2004; Singh and Nair, 2014;
Singh and Chudasama, 2017; 2020; 2021; Ziv et al., 2018). They were expected to ascribe
relationship weights of the sub-concepts (detailed concepts) on a scale of 1–10 as above, and prioritise
the sub-concepts of each main concept separately depending on its relative importance in the system
by assigning the value of 1 to the most important one, and so on. Supplementary material 4 provides
maximum, minimum, and mean values of the relationship strengths of the main concept.

121
Besides holding discussions with IPCC authors on the side-lines of the IPCC meetings, we submitted
requests to 380 authors who had written about certain aspects of the climate system and to
practitioners engaged in adaptation and mitigation measures. In addition, several IPCC authors shared
the instruments with their colleagues and perused them for completion. Through a year-round data
collection effort involving intense persuasion, we ended up collecting 137 responses for instrument
‘a’, 97 for instrument ‘b’, and 148 for instrument ‘c’. On the basis of the IPCC authors’ responses,
additional sub-concepts were introduced in ‘b’ and ‘c’, during the advanced stage of the data
collection process, some of the earlier respondents updated their answers. The participants hailed from
all six of the world’s continents while a quarter of the respondents were female.

3.2. Constructing adjacency weight matrix

Individual cognitive maps of both main and detailed (sub) concepts were coded into square adjacency
weight matrices separately, where the concepts were listed on the vertical and horizontal axes on a
spreadsheet forming an adjacency weight matrix (Singh and Chudasama, 2017; 2020; 2021; Ziv et al.,
2018; Fonseca et al., 2020). The values allocated by the participants were normalised between –1 and
+1 before being encoded into an adjacency weight matrix, where −1 indicates a strong negative
relationship and +1 indicates a strong positive relationship (Özesmi and Özesmi, 2004; Singh and
Chudasama, 2017; 2020; 2021; Ziv et al., 2018; Fonseca et al., 2020; Pereira et al., 2020). The weight
values assigned by the participants to each link between the concepts were coded into the adjacency
matrix (ibid).

3.3. Quantitative aggregation of individual cognitive maps

Individual cognitive maps were quantitatively aggregated to construct a social cognitive map (SCM)
for both main and sub-concepts using the matrix addition (Özesmi and Özesmi, 2004; Singh and
Chudasama, 2017; 2020; 2021; Gray et al., 2019). Most studies aggregate the detailed weight matrix
qualitatively to arrive at a condensed weight matrix with fewer concepts. Since we had elicited
knowledge for both the main and sub-concepts independently, this approach carried the advantage of
triangulation. An arithmetic mean of the weight values was computed to comprehend the cumulative
perception of all the study participants (Özesmi and Özesmi, 2004; Singh and Chudasama, 2017;
2020; 2021; Ziv et al., 2018; Morone et al., 2019). The SCMs derived from the main and detailed
concepts of weight matrices reflected the perceptions of the study participants.

3.4. Condensation of social cognitive map

While we already had the main concepts in a condensed form, the SCM obtained from the quantitative
aggregation of the sub-concepts was further aggregated by the arithmetic mean of the links’ weight
values in each of the main concept categories (Singh and Chudasama, 2017; 2021; Ziv et al., 2018).
The resultant concept weight matrix had the same number (36) of regular/inter-concept links with
seven additional self-loops (Supplementary material 5).

3.5. Preparing final prioritisation score

We calculated the mode values of the prioritisation score to arrive at a final prioritisation of the sub-
concepts on the basis of their relative importance, with the most important sub-concept comprising the
lowest value at 1, and so on. Where more than one mode value was found for a specific sub-concept,
we prioritised the sub-concept with the highest number of the second-lowest prioritisation values over
another sub-concept.

122
3.6. Structural analysis

The matrix algebra of graph theory helps examine the structure of cognitive maps through several
indices, notably density, in-degree, out-degree, degree centrality, complexity, and so on (Özesmi and
Özesmi, 2004; Diestel, 2016; Gray et al., 2019; Morone et al., 2019). The in-degree shows the
cumulative strength of links entering the concept while the out-degree represents the collective
strengths of links exiting the concept. The summation of in-degree and out-degree represents the
degree centrality of a concept. The higher the value of degree centrality, the greater the importance of
a concept in the system (Özesmi and Özesmi, 2004; Diestel, 2016). The density index, a product of
the number of concepts and links, represents the connectivity of a cognitive map (Singh and
Chudasama, 2020). Supplementary materials 6, 7 and 8 provide the structural analysis of the SCM of
the main, condensed, and detailed concepts matrices respectively.

3.7. FCM-based simulations

FCM-based simulations give a deeper understanding of the systems’ behaviour, and the extent to
which one concept impacts the others (Özesmi and Özesmi, 2004; Nápoles et al., 2016; Morone et al.,
2019; Singh and Chudasama, 2020; 2021). Thus, the FCM-based simulations help identify the critical
role and relative importance of a concept. The FCMs can be represented using four functions: FCMs =
(C, M, A, f): (i) C is the total number of concepts in the map, (ii) M : (Ci, Cj) → wij represents the
square adjacency weight matrix of the cognitive map, (iii) Mn×n, A : (Ci) → A(t)i is a state vector that
computes the activation degree of each concept Ci at the discrete-time step t (t = 1, 2, . . . , T), and (iv)
f (x) is the transformation function (Nápoles et al., 2016; Gray et al., 2019; Pereira et al., 2020; Singh
and Chudasama, 2020; 2021).

FCM-based simulation results have been generated by multiplying the initial-state vector value Ai of a
concept Ci with the adjacency weight matrix M. The activation rule and the transformation function
decide the outcome of simulations (Nápoles et al., 2016; Singh and Chudasama, 2017; 2020; 2021;
Gray et al., 2019; Pereira et al., 2020). The activation rule computes the activation vector using the
initial-state vector value of the concepts (ibid). We used the activation rule introduced by Stylios et al.
(1997), as shown in equation (1), because of its memory capabilities.

𝑛
(𝑡+1) (𝑡) (𝑡)
𝐴𝑖 =𝑓 𝐴𝑖 + ∑ 𝑤𝑖𝑗 × 𝐴𝑗 (1)
𝑗=𝑖
𝑗=1
( )

Here, Ai(t+1) is the vector value of concept Ci at the simulation step t+1; Ai(t) is the vector value of
concept Ci at the simulation step t; n is the total number of concepts, while wij is the weight of the link
between concept Ci and concept Cj. Aj(t) is the vector value of concept Cj at the simulation step t while
𝑓( ) stands for the type transformation function deployed during the simulation process.

The concepts are mapped to the real-valued activation level where Ci takes values in the interval [0,1],
which is the degree to which the observation belongs to the concept. Because of the dynamic
influence of connected nodes, the concept’s state changes over time (Nápoles et al., 2016; Pereira et
al., 2020; Singh and Chudasama, 2020; 2021).

We used the hyperbolic tangent transformation function, as shown in equation (2) because the state
vector values of concepts are both positive and negative (Singh and Chudasama, 2017; 2020; 2021;
Pereira et al., 2020).

123
𝑒 𝜆𝑥 − 𝑒 −𝜆𝑥
𝑓 (𝑥 ) = , 𝑎 ∈ R+ (2)
𝑒 𝜆𝑥 + 𝑒 −𝜆𝑥
(t)
Here 𝑥 is the vector value Ai of a concept for a given iteration; e is a residual that describes the
minimum error difference among the subsequent concepts; 𝛌 determines the steepness of the slope;
and R+ is used to keep the activation value of concepts in the (–1, +1) interval (Carvalho, 2013).

3.8. Development of input vectors for scenarios

We used the SCMs of the main and condensed weight matrices, along with the detailed concepts’
(sub-concept) weight matrix for the simulations. The simulation results of the main and condensed
weight matrices provided us the impacts of climate change and the critical roles of mitigation,
adaptation, SDGs, and enabling conditions, while the same for the detailed concepts weight matrix
was used to identify the role of various enabling conditions in promoting SDGs, climate action, and
reduction of climate risks.

We developed eight (baseline + seven) scenarios for FCM-based simulations of the main and
condensed weight matrices to identify various climate-resilient development pathways. The baseline
scenario was implemented by activating the concept C3: climate change—while considering the
increasing climate change trends following global projections. Identifying critical concepts for
scenario preparation helps connect storylines to the quantitative model (Amer et al., 2016). We
identified individual concepts from the main and condensed weight matrices and their various
plausible combinations, depending on their mutual compatibility (i.e., input vectors) in order to
develop climate-resilient development pathways. The seven input vectors corresponding with seven
independent scenarios were: (i) C7: shocks; (ii) C9: mitigation; (iii) C8: adaptation; (iv) C10:
sustainable development; (v) C8: adaptation, C9: mitigation, and C10: sustainable development; (vi)
C8: adaptation, C9: mitigation, C10: sustainable development, and C11: enabling conditions; and (vii)
C7: shocks, C8: adaptation, C9: mitigation, C10: sustainable development, and C11: enabling
conditions. The concept C3: climate change was also activated with every input vector.

On the basis of results of the main and condensed weight matrices, which established the critical role
of enabling conditions, we conducted eleven scenarios utilising the detailed concepts weight matrix to
identify the relative roles of various enabling conditions — one baseline scenario by activating all the
sub-concepts of climate change, and ten scenarios for individual enabling conditions. Input vectors,
corresponding to ten independent scenarios, were arranged on the basis of prioritisation ranking
assigned by the study participants. These included: (i) EC3: Dimensions of governance (policy,
institutions, practice); (ii) EC4: Actors and arena of engagements; (iii) EC7: Ethics values and
worldviews; (iv) EC2: Societal transformation; (v) EC9: Innovations; (vi) EC5: Partnership &
commitment to finance and technology; (vii) EC1: Social justice; (viii) EC10: Digital transformation;
(ix) EC6: Feasibility; and (x) EC8: Regenerative and conscious capitalism.

A baseline was created to initiate the simulation process by activating the initial value of the climate
change concept, in both instances, consistent with the long-term climate change projections. The
simulation outcomes were then compared against the baseline. The climate change concept was
‘activated’ every time for every scenario, in view of climate change. Exploring the changes in concept
values between the baseline and other scenarios allows for a quantitative interpretation of the impact
of the key concepts on the system in relative terms.

3.9. Simulation Process

124
Each concept in the FCM system has an input vector value that varies from |0| to |1|, corresponding
to a ‘non-activated’ or ‘activated’ state, in that order (Pereira et al., 2020; Singh and Chudasama,
2020, 2021). The ‘activated’ concepts influence other concepts following the weighted cause-and-
effect relationships of the adjacency matrix (ibid). Due to the recurrent neural network nature of FCM,
the simulation process is iterated by multiplying the ‘activated’ concepts (i.e., input vector) with the
adjacency matrix and by applying the threshold squashing function (equation 2) after each
multiplication until the FCM system converges to a steady-state (Carvalho, 2013; Nápoles et al.,
2016; Singh and Chudasama, 2017; 2020; 2021; Morone et al., 2019; Pereira et al., 2020). As both the
main and detailed FCM systems have multiple feedback loops, repeated activation of concepts sets in
as simulations occur (Nápoles et al., 2016; Singh and Chudasama, 2017; 2020; 2021; Morone et al.,
2019). We implemented the simulation process in each scenario described in section 3.8 with the
initial state vector of the input vectors clamped to 1 (A1) and the initial state vector of all the other
concepts clamped to 0 (A0). The FCM-based simulation process was implemented using the
FCMWizard. We also ascertained the sensitivity of the system by clamping the concepts of each input
vector to other non-zero values namely, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, and 0.9 to determine the
stability of the system in yielding simulation results.

4. Results

Enhancing planetary health and human well-being is the main aim of climate resilient development
pathways. Multiple factors play a critical role, directly or indirectly, in achieving this aim. Certain
variables including socio-economic drivers, climate processes, climate change, shocks, consequent
adverse impacts, and losses and damages, along with future threats wield a negative impact on CRD.
On the other hand, variables like adaptations and mitigation efforts, along with achieving sustainable
development goals (SDGs), and their enabling conditions help advance planetary health and human
well-being. Multiple interconnections and feedback loops characterise the above-mentioned
dimensions (Fig. 2).

4.1. Structural analysis

Various dimensions of the network statistics of the three weight matrices included in the study —
main, condensed, and detailed — are contained in supplementary materials 6, 7, and 8. The SCM of
the main concepts weight matrix comprises 12 concepts and 36 links (Supplementary material 6). The
contribution of a concept in a cognitive map may be understood by its degree of centrality- the
summation of in-degree and out-degree. The matrix has a high-density index of 0.25, which implies
that 25% of links are present relative to the maximum number of links that may potentially occur
between the 12 concepts. The SCM of condensed concepts weight matrix comprises 12 concepts and
43 links, of which 36 connections are regular and seven self-loops (Supplementary material 7). The
matrix has an extremely high-density index of 0.298. Although the simulation results of both the main
and condensed concept weight matrices (with similar inter-concept links) reflect similar findings
(Table 1; Supplementary material 9), most values of the condensed concept weight matrix normalised
due to seven additional self-loops in the condensed weight matrix. The detailed concepts weight
matrix represents an extremely complex system with 107 concepts and 360 connections
(Supplementary material 8). Few studies have used such a large number of concepts available in our
detailed weight matrix for simulations.

4.2. Prioritisation ranking of important sub-concepts

125
Respondents prioritised sustainable agriculture, sustainable water management, climate-smart
agriculture, and conservation agriculture as the most important adaptation measures (Fig. 2). They
accorded high importance to carbon dioxide removal, industrial system transition, energy system
transition, and land and ecosystem transition under the mitigation category (Fig. 2). The respondents
accorded high importance to SDGs 1 (end poverty), 2 (food and nutritional security; sustainable
agriculture), 4 (equitable quality education), 13 (Climate action), and 7 (Affordable and sustainable
modern energy) for implementation (Fig. 2). Of the enabling conditions, the respondents accorded
high importance to the facets of governance, actors and arenas of engagement, along with ethics,
values, and worldviews in shaping the pathways for CRD (Fig. 2).

4.3. Perturbation of the socio-ecological system

The key drivers of climate change and past developments, including maldevelopment trends, lead to
climate processes and eventually to climate variability and change characterised by increasing
temperatures and global warming, rising sea-levels, oceanic acidification, precipitation variabilities,
and weather extremes including droughts, heatwaves, floods, and cyclones (Fig. 2).

The FCM-based simulation of the main/condensed weight matrix suggests that climate change will
adversely impact the socio-economic and ecological systems. Climate change compound adverse
impacts, and losses and damages, while enhancing potential risks to socio-economic and ecological
systems. Hence, a dangerously warm planet will become extremely inhospitable for human
civilisation (Table 1: baseline scenario/ Supplementary materials 9 and 10: baseline scenario).

Our FCM-based SCM simulations of the detailed concept weight matrix suggest that increased
warming and the resultant impacts (baseline scenario) could accelerate water insecurity, land
degradation, decrease productivity, increase food and fodder insecurity, reduce ecosystem
productivity, and impact human health while affecting natural, ecological, and human systems
(Supplementary materials 11a and 12a). The baseline scenario also indicates a range of potential
threats and damages including loss of habitat, biodiversity, human lives, and economic loss, along
with damages to infrastructure, settlements, and cultural heritage, among other things (Supplementary
materials 11b and 12b). According to the baseline scenario, there are potential threats of water and
food shortages, climatic extremes (droughts, floods, heatwaves, etc.), decreased living standards and
equity, and reduced human health. Climate change could also impact ecosystems and their services
along with critical physical infrastructure and networks while threatening peace and human
settlements (Supplementary materials 11c and 12c). Overall, the baseline scenario indicates that
climate change will render our planet unlivable.

4.4. Impacts of shocks

The FCM-based simulations of the main/condensed concept SCM demonstrate that the pathways of
CRD are likely to be turbulent because of large-scale external shocks like COVID-19, with synthetic
biology generating bacteria and viruses, nuclear wars, and solar radiation management (Table 1:
scenario 1; Supplementary material 9 and 10: scenario 1). Climate change, along with large-scale
external shocks, is capable of aggravating key risks, climate processes, and climate change (Table 1:
scenario 1; Supplementary materials 9 and 10: scenario 1). Nuclear warfare and solar radiation, too,
can wield an adverse influence on climate change.

126
Fig. 2: Dimensions for modelling climate resilient development
Note: Values in columns ‘P’ shows ‘mode values’ of prioritisation by the research participants with 1 representing as the most important

127
4.5. Climate action and SDGs in attaining CRD

Our findings of the FCM-based simulations of the main/condensed concepts SCM demonstrate that
climate adaptation or mitigation actions alone cannot restore ecosystems (Table 1: scenario 2 and 3;
Supplementary materials 9 and 10: scenario 2 and 3). Measures of SDGs, and more importantly
adaptation and mitigation measures along with meeting SDG targets, can effectively reduce impacts,
and losses and damages to make way for climate-resilient development (Table 1: scenario 4 and 5;
Supplementary materials 9 and 10: scenario 4 and 5; Fig. 3).

Adaptation, mitigation and SDG measures, coupled with the right enabling conditions, could reduce
climate change, its impacts, losses and damages, and future risks to a greater degree; they will be
required for robust climate-resilient development (Table 1: scenario 6; Supplementary materials 9 and
10: scenario 6; Fig. 3). Results indicate that these combinations can provide climate-resilience even in
the face of shocks (Table 1: scenario 7; Supplementary materials 9 and 10: scenario 7). Climate action
and sustainable development are interdependent and need to be pursued in an integrated manner to
enhance human well-being on a healthy planet.

4.6. Enabling conditions of climate resilient development

Our FCM-based simulations of the main SCM established the critical role of enabling conditions in
promoting the successful implementation of adaptation and mitigation, resilience building, and SDG
achievement (Table 1: scenario 6; Supplementary materials 9 and 10: scenario 6; Fig. 3), while
significantly reducing the negative impacts of shocks (Table 1: scenario 7; Supplementary materials 9
and 10: scenario 7; Fig. 3). In order to establish the relative role of enabling conditions in meeting
SDGs and implementation of adaptation and mitigation actions, reducing key risks, and ultimately
achieving planetary health and human well-being, we implemented simulation exercises using the
detailed SCM.

Simulation analysis of the detailed SCM revealed a somewhat different image to the prioritisations
established by the study participants. Considering the strength of the FCM model demonstrating
causal relationships and feedback loops in a complex system, the FCM-based simulations could help
identify the critical role of specific concepts with rigor while offering valuable insights for
policymaking. Human beings are unlikely to visualise such complex causal interactions occurring in a
complex system.

The FCM-based simulations of the detailed (sub) concepts reveal similar findings on the role of
enabling conditions in reducing the impacts of climate change, reducing losses and damages, and
managing the future consistent with the main concepts. The FCM-based simulations of the detailed
(sub) concepts further reveal that enablers like: (i) ethics, values, and worldviews, (ii) partnerships
and commitment to finance and technology by the national governments, (iii) actors and arena of
engagements, (iv) innovations, and (v) dimensions of governance that are particularly critical to
facilitating the implementation of SDGs, mitigation, and adaptation actions (Supplementary materials
11d to 11f and 12d to 12f).

128
Table 1: Simulation results of the main SCM

(IV stands for initial value; FV stands for final value)


Baseline Scenario 1 Scenario 2 Scenario 3 Scenario 4 Scenario 5 Scenario 6 Scenario 7
Concepts
IV FV IV FV IV FV IV FV IV FV IV FV IV FV IV FV
C1: Key drivers 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
C2: Climate processes 0 0.9033 0 0.9217 0 0.9033 0 0.9032 0 –0.9030 0 –0.9032 0 –0.9741 0 –0.9741
C3: Climate change 1 0.9208 1 0.9233 1 0.9208 1 0.9207 1 –0.9206 1 –0.9207 1 –0.9297 1 –0.9297
C4: Adverse impacts 0 0.9805 0 0.9841 0 0.9805 0 0.9805 0 –0.9804 0 –0.9805 0 –0.9947 0 –0.9947
C5: Losses and Damages 0 0.9986 0 0.9989 0 0.9986 0 0.9986 0 –0.9986 0 –0.9986 0 –0.9997 0 –0.9997
C6: Key risks 0 0.9801 0 0.9840 0 0.9801 0 0.9801 0 –0.9801 0 –0.9801 0 –0.9986 0 –0.9986
C7: Shocks 0 0 1 0.1433 0 0 0 0 0 0 0 0 0 –0.8924 1 –0.8924
C8: Adaptation 0 –0.9264 0 –0.9265 0 –0.9264 1 –0.9264 0 0.9263 1 0.9264 1 0.9861 1 0.9861
C9: Mitigation 0 –0.9268 0 –0.9269 1 –0.9268 0 –0.9268 0 0.9267 1 0.9267 1 0.9870 1 0.9870
C10: Sustainable development 0 –0.9960 0 –0.9967 0 –0.9960 0 –0.9960 1 0.9960 1 0.9960 1 0.9998 1 0.9998
C11: Enabling conditions 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1
C12: Planetary health and human well-being 0 –1 0 –1 0 –1 0 –1 0 0.9999 0 0.9999 0 1 0 1

129
Fig. 3: Pathways for climate resilient development
Note: This figure is based on FCM-based simulations of main and detailed SCMs. The dimensions of enabling conditions are sorted based on their role in facilitating SDGs,
mitigation and adaptations with the most crucial one at the top.

130
Thus, FCM-based simulations laid out the functions of enabling conditions to promote the successful
implementation of SDGs and climate actions to advance the achievement of planetary health and
human well-being in the following order of decreasing importance: (i) ethics, values, and worldviews;
(ii) partnerships and commitment to finance and technology by the national governments; (iii) actors
and arena of engagements; (iv) innovations; (v) dimensions of governance; (vi) social justice; (vii)
digital transformations; (viii) societal transformation; (ix) economic, financial, technological, social
and geopolitical feasibilities of system transitions; and (x) regenerative and conscious capitalism
(Supplementary materials 11a to 11f and 12a to 12f). In order to advance planetary health and well-
being for everyone, the enabling conditions need to function in tandem.

5. Discussion: Charting CRDPs

Charting climate resilient development pathways (CRDPs) entails facilitating the process of climate
change adaptation, mitigation and sustainable development. Climate change mitigation and adaptation
actions primarily help reduce GHG emission concentration, atmospheric aerosols loading, climate
change impacts, losses and damages, and climate change risks (IPCC, 2019; Allen et al., 2018;
O'Neill et al., 2017). Coordinating adaptation and mitigation measures with SDGs could help
streamline funds and efforts to prevent inconsistencies and accomplish ‘triple-wins’ (Nunan, 2017).

Ethics, values, and worldviews shape the narratives for the engagement of actors (Omukuti, 2020)
across multiple arenas (Fig. 4). Ethics (dignity, equity, compassion) and values often constitute
leveraging points for shifting unsustainable development trajectories (O’Neill et al., 2015; O’Brien,
2018). Public support for global climate policies and public response to climate change in the domain
of adaptation and lifestyle changes are closely related to and are articulated via people’s perceptions,
ethical concerns, values, and worldviews (Hayward, 2012; Wolf et al., 2013).
Building climate change resilience requires substantial financial investments and changes in political
and institutional foundations supporting adaptation and mitigation actions and sustainable socio-
economic development (Allen et al., 2018; de Coninck et al., 2018; Roy et al., 2018). Article 9.2 of
the Paris Agreement (UNFCCC, 2015) reiterates the requirement for developed countries to provide
financial support to developing countries for both mitigation and adaptation with developing
countries, possibly, voluntarily contributing to the financing efforts. Climate financing included in the
US $100 billion annual pledge under the Paris Agreement is insufficient for climate change mitigation
and adaptation initiatives (Peake and Ekins, 2016). There has been, however, a significant difference
between the pledge and the actual distribution of funds (Khan et al., 2020). Article 9.2 of the Paris
Agreement (UNFCCC, 2015) links adaptation needs to the degree of mitigation. While adaptation
financing stands at about 20% of global climate finance, the US $ 50 billion pledge is not enough to
meet adaptation needs worldwide. Adaptation needs are likely to go up to US $ 150–300 by 2030 and
US $ 300–500 by 2050 (UNEP, 2016). Adaptation financing has remained limited relative to
mitigation, accounting for 5-7% of the overall monitored climate finance at US $100 billion per
annum since 2015 (Carty et al., 2020). For instance, in 2018, developing countries received less than
25% of the total finance, 66% of which went into mitigation and 9% to cross-cutting projects (ibid).
Renewed levels of public finance commitments and increased flow of climate finance from developed
countries and budgetary allocations of climate finance by developing countries are imperative even in
the short-term (2021–2030). Besides, low-income developing countries require substantially high
development financing, including official development assistance for SDGs (Gaspar et al., 2019).
Similarly, technology transfer from developed to developing countries can promote dematerialisation
and decarbonisation measures (Hansen et al., 2019) while adding the achievement of sustainable
production and consumption.

131
Fig. 4: Enabling conditions facilitating CRD

Note: Interactions between actors and arena of engagements across various scales are extremely complex and
shaped by ethics, values and worldviews

Dynamic interactions between governance actors, modes, and dimensions of governance [policy,
institutions (norms, rules), and practice] (Sørensen, 2017) constitute arenas of engagement across
domains of action and decision-making—political, economic, ecological, socio-cultural, community,
and knowledge and technology that could eventually shape CRDPs (Fig. 4). Multiple actors —
government, industry, science, media, and civil society — have been engaging across multiple arenas
with asymmetric power relations to shape climate narratives, actions and development goals. Such
engagements occur across various scales — local, sub-national, national, regional, and global — and

132
are bound by the ethics, values and worldview of actors (Fig. 4). CRD will require multi-level, multi-
stakeholder and transnational governance arrangements involving a variety of state and non-state
actors participating by varying capabilities and authority. However, a need for an enforceability
mechanism for effective accountability of such networks (Banerjee, 2010) and compliance-related
norms and soft laws for ensuring accountability (Korbin, 2009) cannot be ignored. New networked
climate governance models have established multi-scalar polycentric governance structures (Jordan et
al., 2015) that still require a defined legal framework. (Mansbridge, 2014). Ostrom (2010) also
predicted that polycentric governance would be more diverse, multi-scalar, with a greater focus on
bottom-up measures to meet the needs of different actors.

The SR1.5 states that ‘without societal transformation and rapid implementation of ambitious
greenhouse gas reduction measures, pathways to limiting warming to 1.5°C and achieving sustainable
development will be exceedingly difficult, if not impossible, to achieve’ (Roy et al., 2018). The
societal transformations are deeply political, involving power struggles and value conflicts (Patterson
et al., 2016). Such a transformation requires the use of innovative technologies, inventive use of
markets, and deeper structural change (Scoones et al., 2015). However, without integrating social
justice, the efforts of technology-led, market-led, or state-led green transitions towards a low-carbon
world, are unlikely to succeed and remain sustainable in the long term (Scoones et al., 2015). Besides,
social justice and equity as core aspects of CRD demand equality. As Sen (2009) suggested, a theory
of justice must provide both the fairness of the process and the fair distribution of opportunities.

Societal transformations call for accelerated innovations of all kinds— technological, economic,
institutional, and social. Innovations can expand in scale, scope, and geography and ultimately help
generate new potential pathways for CRD. The digital revolution can help enhance productivity while
lowering production costs, reducing emissions, improving resource efficiencies, supporting the
circular economy, enabling zero-carbon energy systems, and monitoring ecosystems (Sachs et al.,
2019) that are critical to achieving CRD.

The SR1.5 identifies four simultaneous and necessary system transitions to enable adaptations and
mitigation actions— energy, industrial, land and ecosystems, urban and infrastructure (de Coninck et
al., 2018). System transitions are not a neutral process and have implications for social justice and the
long-term ecological consequences of proposed solutions (Blythe et al., 2018). These system
transitions can only be accomplished when their economic, technological, institutional, socio-cultural,
environmental, and geophysical feasibilities are ascertained (ibid). The system transitions can not only
widen the solution space but also accelerate and deepen the process of CRD. However, such
transitions may pose potential risks to certain sectors, regions, or societies. Hence, managing
transition risk is a critical element of transforming society.
Systemic corporate strategies of regenerative and conscious capitalism could help industries create a
self-organising, self-maintaining, highly adaptive, and regenerative form of capitalism that generates
lasting social and economic vitality and human well-being (Fullerton, 2015). These principles could
be enabling factors of the industrial system transition for a dematerialised and decarbonised world.

Today, the issue is whether the world will have just one pathway or whether each country will create
its own. Although both planetary health and human well-being are profoundly ethical issues, planetary
health is perturbed by ongoing biogeochemical flows, climate change and resultant impacts on
biosphere integrity, and land system change (Steffen et al., 2015). Human well-being, though, is more
locale-specific. Low and middle-income countries have the added responsibility of addressing the
basic needs (poverty, hunger, nutritional security, access to safe drinking water, sanitation and energy,

133
and inclusive development) and survival needs (physiology, safety, security) of their citizens. In
developed countries where basic needs are mostly satisfied, some survival needs (safety, security),
particularly psychological needs (love, belonging, esteem, self-actualisation), and spiritual needs
(self-transcendence) need to be addressed. Besides, social equity and social justice are other issues the
world needs to pay heed to. Affluent-apposite pathways may impose notable impacts and costs on the
vulnerable sections of the global society (Hickel, 2017), directly contradicting the commitment to
‘leave-no-one-behind’ in meeting the SDGs (UN, 2015). The imminent and pervasive challenge is to
identify ways in which the well-off can live within a safe limit of planetary boundaries while allowing
the poor to lead decent lives (Hayward, 2012). Hence, each nation may have its own pathways for
human well-being and eventually CRDPs, depending on its current stage of development, and pace of
SDG progress.
One may also ask: on what time scale should the CRD be defined? As Costanza and Patten (1995)
stated, ‘nothing lasts forever, not even the universe as a whole.’ Kates (2011) proposed sustainability
issues over the span of a century. The timeline for CRD by the end of the twenty-first century seems
plausible considering the complexities of planetary-scale phenomena, the high residence time of the
greenhouse gases in the environment, technological breakthroughs, shifts in human worldviews, and
the subsequent interactions of facilitating circumstances.
There are no established pathways that will fulfill all the SDGs and the goals of the Paris Agreement.
Mediated by ethics, values and worldviews, the CRD will be the outcome of a sequence of decisions,
actions, and interventions undertaken by different actors under different arenas of engagements at
local, sub-national, national, regional and global scales for shaping development trajectories that
eventually influence the state of planetary health and human well-being. Based on the continuum of
interactions between enabling conditions for adaptation, mitigation, and SDG actions in the short term
(2021–2030), dimensions of sustainable development in medium to long term (beyond 2030),
different pathways for planetary health and human well-being will evolve for the developed and
developing societies (Fig. 5). Thus, a continuum of planetary health and human well-being will evolve
centered on pathways emerging as a vector product of the interactions between enabling conditions
for implementation of adaptation, mitigation, SDGs/ sustainable development dimensions. While low
planetary health represents a dangerously warm world, disrupting ecosystem processes beyond its
regenerative potential and impairing ecosystem services, high planetary health suggests restricting
climate change and, maintaining regenerative ecosystem processes to ensure that the ecosystem
provides its provisioning, regulating, supporting and cultural services adequately. While low human
well-being signifies inequality, injustice, alienation from nature and society, vulnerability,
dissatisfaction, disesteem, displeasure, self-degradation, and subservience, high human well-being
signifies equity, justice, connectedness to nature and society, resilience, fulfillment, esteem, ultimate
pleasure, self-actualisation, and self-transcendence. Consequently, each nation may have its own
pathways of climate resilient development emerging as the vector product of the interactions between
enabling conditions, its current stage of development, and pace of adaptation, mitigation and SDG
progress. Accordingly, there will also be a continuum of climate resilient development in place (Fig.
5). Thus, CRD is not only an outcome but also a process of actions and social choices made by
multiple actors. A holistic worldview by all actors that promotes co-productive balance between
human and nature, while providing healthy and meaningful livelihoods for all through the just and
equitable sharing of resources, is likely to enhance CRD.

134
Fig. 5: Evolving pathways for climate resilient development

Note: Climate resilient development pathways emerges as a vector product of the interactions between enabling
conditions for adaptation, mitigation, and development actions

135
6. Conclusions
Achieving planetary health and human well-being requires synergies in the implementation of
mitigation, adaptation and SDG actions. Our FCM-based simulations demonstrated the critical roles
of the following enabling conditions in advancing CRD: (i) ethics values and worldviews; (ii)
partnerships and commitment to finance and technology by the national governments; (iii) actors and
arena of engagements across local to global scales; and (iv) innovations.
CRD transformation requires complementary and co-produced actions by multiple actors. The
pathways for CRD involve struggles between divergent worldviews articulated in political, socio-
cultural, economic, knowledge-technology, ecological, and community discourses. It is important to
strengthen transnational and trans-local networked governance structures with an appropriate legal
framework under polycentric arrangements for crafting and legitimising societal choices to advance
planetary health and human well-being.

Accordingly, we define climate resilient development as a development embracing mitigation,


adaptation and inclusive sustainable development involving deep societal transformation to advance
planetary health and well-being for all. Thus, CRD enables humankind and nature to thrive in unison.

Bounded by ethics, values, and worldviews, the emerging continuum of complex interactions between
various enabling conditions at local, sub-national, national, regional and global scales will promote
the degree of adaptation, mitigation, and SDG achievements, thus dynamically evolving the level of
human well-being on a healthy planet along the continuum (Fig. 5). As a consequence, the CRDP will
be an evolving resultant vector product of complex interactions occurring between various enabling
conditions to facilitate SDGs and climate actions.
CRD is more of a journey than a destination. ‘Every long-term pathway depends crucially not just on
actions by today’s decision-makers, but also by future decision-makers and future generations’
(Pörtner et al., 2014), However, short-term decisions and actions will have huge implications for long-
term impacts in shaping the pathways for CRD. Besides, the window of opportunity for a decent life
for globally vulnerable communities will dramatically reduce in the future if the right actions are not
taken now. Changes in actors’ ethics, beliefs and worldviews will help in re-engineering climate and
development narratives and actions, and hence the potential direction of CRD. Undoubtedly, the role
of ethics, values, worldviews, and social justice will be critical to the deep societal transformation and
protection of citizens against climate change as a human right. All developed and developing societies
need to chart their context-specific pathways to CRD, bearing in mind that a shared planet needs to
regenerate healthier and meaningful livelihoods through the just and equitable sharing of resources
across the planet. The human race needs to follow a value-driven system of production and
consumption in both industrialised and developing nations that value biodiversity, reintegrates
humans with nature, and allows us to flourish as genuinely intelligent, sensible, and judicious beings
in the Anthropocene.

136
References

Allen, M., Dube, O.P., Solecki, W., Aragón–Durand, F., Cramer, W., Humphreys, S., Kainuma, M., Kala, J., Mahowald, N.,
Mulugetta, Y., Perez, R., Wairiu, M. and Zickfeld, K., 2018. Framing and Context. In: Global Warming of 1.5°C. An
IPCC Special Report on the impacts of global warming of 1.5°C above preindustrial 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 [Masson-Delmotte, V., Zhai, P., Pörtner, H.-O., Roberts, D.,
Skea, J., Shukla, P.R., Pirani, A., Moufouma-Okia, W., Péan, C., Pidcock, R., Connors, S., Matthews, J.B.R., Chen, Y.,
Zhou, X., Gomis, M.I., Lonnoy, E., Maycock, T., Tignor, M. and Waterfield, T. (eds.)]. In press.

Amer, M., Daim, T.U. and Jetter, A., 2016. Technology roadmap through fuzzy cognitive map-based scenarios: the case of
wind energy sector of a developing country. Technology Analysis & Strategic Management, 28(2), pp.131–155.
https://doi.org/10.1080/09537325.2015.1073250.

Banerjee, S.B., 2010. Governing the global corporation: A critical perspective. Business Ethics Quarterly, 20(2), pp.265–
274. https://doi.org/10.5840/beq201020219.

Blythe, J., Silver, J., Evans, L., Armitage, D., Bennett, N.J., Moore, M.-L., Morrison, T.H. and Brown, K., 2018. The Dark
Side of Transformation: Latent Risks in Contemporary Sustainability Discourse. Antipode, 50(5), pp.1206–1223,
https://doi.org/10.1111/anti.12405.

Carty, T., Kowalzig, J. and Zagema, B., 2020. Climate Finance Shadow Report 2020: Assessing progress towards the $100
billion commitment. Oxfam International, Oxford. [online] Available at: https://www.oxfam.org/en/research/climate-
finance-shadow-report-2020 [Accessed 14 November 2020].

Carvalho, J.P., 2013. On the semantics and the use of fuzzy cognitive maps and dynamic cognitive maps in social sciences.
Fuzzy Sets and Systems, 214, pp.6–19. https://doi.org/10.1016/j.fss.2011.12.009.

Conway, D., van Garderen, E.A., Deryng, D., Dorling, S., Krueger, T., Landman, W., Lankford, B., Lebek, K., Osborn, T.,
Ringler, C., Thurlow, J., Zhu, T. and Dalin, C., 2015. Climate and Southern Africa’s water-energy-food nexus. Nature
Climate Change, 5, pp.837–846. https://doi.org/10.1038/nclimate2735.

Costanza, R. and Patten, B.C., 1995. Defining and predicting sustainability. Ecological Economics, 15(3), pp.193–196.

Craft, B. and Fisher, S., 2018. Measuring the adaptation goal in the global stock take of the Paris Agreement. Climate Policy,
18(9), pp.1203–1209. https://doi.org/10.1080/14693062.2018.1485546.

Daly, H.E., 1995. On Wilfred Beckerman’s critique of sustainable development. Environmental Values, 4, pp.49–55.

Dasgupta, P., 2001. Human Well-Being and the Natural Environment. Oxford University Press, p.328. ISBN-
13:9780199247882.

de Coninck, H., Revi, A., Babiker, M., Bertoldi, P., Buckeridge, M., Cartwright, A., Dong, W., Ford, J., Fuss, S., Hourcade,
J.-C., Ley, D., Mechler, R., Newman, P., Revokatova, A., Schultz, S., Steg, L. and Sugiyama, T., 2018. Strengthening
and implementing the global response. In: 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 [Masson-Delmotte, V., Zhai, P., Pörtner, H.-O.,
Roberts, D., Skea, J., Shukla, P.R., Pirani, A., Moufouma-Okia, W., Péan, C., Pidcock, R., Connors, S., Matthews,
J.B.R., Chen, Y., Zhou, X., Gomis, M.I., Lonnoy, E., Maycock, T., Tignor, M. and Waterfield, T. (eds.)]. In press.

de Loma-Osorio, G.F.Y., 2016. The 2030 Agenda for Sustainable Development: Bringing Climate Justice to Climate Action.
Development, 59(3), pp.223–228. https://doi.org/10.1057/s41301-017-0122-9.

Denton, F., Wilbanks, T.J., Abeysinghe, A.C., Burton, I., Gao, Q., Lemos, M.C., Masui, T., O’Brien, K.L. and Warner, K.,
2014. Climate-resilient pathways: adaptation, mitigation, and sustainable development. In: Climate Change 2014:
Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the
Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., Barros, V.R., Dokken, D.J.,
Mach, K.J., Mastrandrea, M.D., Bilir, T.E., Chatterjee, M., Ebi, K.L., Estrada, Y.O., Genova, R.C., Girma, B., Kissel,
E.S., Levy, A.N., MacCracken, S., Mastrandrea, P.R., White, L.L., (eds.)]. Cambridge, United Kingdom and New York,
NY, USA: Cambridge University Press. pp.1101–1131.

137
Diestel, R., 2016. Graph theory. Heidelberg: Springer-Verlag. eISBN:978-3-96134-005-7.

Dominati, E., Patterson, M. and Mackay, A., 2010. A framework for classifying and quantifying the natural capital and eco-
system services of soils. Ecological Economics, 69(9), pp.1858–1868.

Doyal, L. and Gough, I., 1991. Introduction. In: Doyal, L. and Gough, I., 1991. A Theory of Human Need. Palgrave, pp.1–2.

Fankhauser, S. and McDermott, T.K.J., 2016. Climate-resilient development: an introduction. In: Fankhauser, S. and
McDermott, T.K.J., 2016. The Economics of Climate-Resilient Development. Edward Elgar Publishing. p.256.
ISBN:978-1-78536-030-5.

Fonseca, J.P.C., Ferreira, F.A.F., Pereira, L.F., Govindan, K. and Meidute-Kavaliauskiene, I., 2020. Analyzing determinants
of environmental conduct in small and medium-sized enterprises: A sociotechnical approach. Journal of Cleaner
Production, 256, pp.120380. https://doi.org/10.1016/j.jclepro.2020.120380.

Fullerton, J., 2015. Regenerative capitalism - How universal principles and patterns will shape our new economy. Capital
Institute.

Ganglmair-Wooliscroft, A. and Wooliscroft, B., 2019. Well-Being and Everyday Ethical Consumption. Journal of
Happiness Studies, 20, pp.141–163. https://doi.org/10.1007/s10902-017-9944-0.

Gaspar, V., Amaglobeli, D., Garcia-Escribano, M., Prady, D. and Soto, M., 2019. Fiscal Policy and Development: Human,
Social, and Physical Investment for the SDGs. International Monetary Fund. p.45. ISBN:9781484388914.

Geels, F.W., Berkhout, F. and van Vuuren, D.P., 2016. Bridging analytical approaches for low-carbon transitions. Nature
Climate Change, 6, pp.576–583. http://dx.doi.org/10.1038/nclimate2980.

Grassi, G., House, J., Dentener, F., Federici, S., den Elzen, M. and Penman, J., 2017. The key role of forests in meeting
climate targets requires science for credible mitigation. Nature Climate Change, 7(3), pp.220–226.
https://doi.org/10.1038/nclimate3227.

Gray, S.R.J., O'Mahony, C., O'Dwyer, B., Gray, S.A. and Gault, J., 2019. Caught by the fuzz: Using FCM to prevent coastal
adaptation stakeholders from fleeing the scene. Marine Policy, 109, pp.103688.
https://doi.org/10.1016/j.marpol.2019.103688.

Griscom, B.W., Adams, J., Ellis, P.W., Houghton Richard A., Lomax Guy, Miteva Daniela A., Schlesinger William H.,
Shoch David, Siikamäki Juha V., Smith P., Woodbury P., Zganjar C., Blackman A., Campari J., Conant R.T., Delgado
C., Elias P., Gopalakrishna T., Hamsik M.R., Herrero M., Kiesecker J., Landis E., Laestadius L., Leavitt S.M.,
Minnemeyer S., Polasky S., Potapov P., Putz F.E., Sanderman J., Silvius M., Wollenberg E. and Fargione J., 2017.
Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), pp.11645–11650.
https://doi.org/10.1073/pnas.1710465114.

Haines, A., Amann, M., Borgford-Parnell, N., Leonard, S., Kuylenstierna, J. and Shindell, D., 2017. Short-lived climate
pollutant mitigation and the Sustainable Development Goals. Nature Climate Change, 7, pp.863–869.
https://doi.org/10.1038/s41558-017-0012-x.

Hansen, E., Lüdeke-Freund, F., Quan, X. and West, J., 2019. Cross-National Complementarity of Technology Push, Demand
Pull, and Manufacturing Push Policies: The Case of Photovoltaics. Institute of Electrical and Electronics Engineers,
66(3), pp.381–397. https://doi.org/10.1109/TEM.2018.2833878.

Hayward, T., 2012. Climate change and ethics. Nature Climate Change, 2, pp.843–848.
https://doi.org/10.1038/nclimate1615.

Hickel, J., 2017. Is global inequality getting better or worse? A critique of the World Bank’s convergence narrative, Third
World Quarterly, 38(10), pp.2208–2222. https://doi.org/10.1080/01436597.2017.1333414.

Horton, R, Beaglehole, R., Bonita, R., Raeburn, J., McKee, M. and Wall, S., 2014. From public to planetary health: a
manifesto. The Lancet, 383(847). https://doi.org/10.1016/S0140-6736(14)60409-8.

Hulme, M., 2018. ‘Gaps’ in Climate Change Knowledge: Do They Exist? Can They Be Filled? Environmental Humanities,
10(1), pp.330–337. https://doi.org/10.1215/22011919-4385599.

138
IPCC, 2019. 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 [Shukla, P.R., Skea,
J., Calvo Buendia, E., Masson-Delmotte, V., Pörtner, H.-O., Roberts, D.C., Zhai, P., Slade, R., Connors, S., van
Diemen, R., Ferrat, M., Haughey, E., Luz, S., Neogi, S., Pathak, M., Petzold, J., Pereira, J.P., Vyas, P., Huntley, E.,
Kissick, K., Belkacemi, M., Malley, J., (eds.)]. In press.

Jordan, A., Huitema, D., Hildén, M., van Asselt, H., Rayner, T.J., Schoenefeld, J.J., Tosun, J., Forster, J. and Boasson, E.L.,
2015. Emergence of polycentric climate governance and its future prospects. Nature Climate Change, 5, pp.977–982.
https://doi.org/10.1038/nclimate2725.

Kates, R.W., 2011. What kind of a science is sustainability science? Proceedings National Academy of Science, 108,
pp.19449–19450. https://doi.org/10.1073/pnas.1116097108.

Khan, M., Robinson, S.-a., Weikmans, R., Ciplet, D. and Roberts, J.T., 2020. Twenty-five years of adaptation finance
through a climate justice lens. Climatic Change, 161, pp.251–269. https://doi.org/10.1007/s10584-019-02563-x.

Klenk, N., Fiume, A., Meehan, K. and Gibbes, C., 2017. Local knowledge in climate adaptation research: moving
knowledge frameworks from extraction to co-production. WIREs Climate Change, 8(5), pp.e475.
https://doi.org/10.1002/wcc.475.

Korbin, S.J., 2009. Private Political Authority and Public Responsibility: Transnational Politics, Transnational Firms, and
Human Rights. Business Ethics Quarterly, 19(3), pp.349–74. https://doi.org/10.5840/beq200919321.

Kristensen, H.M. and Korda, M., 2020. Status of World Nuclear Forces, Federation of American Scientists. [online]
Available at: https://fas.org/issues/nuclear-weapons/status-world-nuclear-forces [Accessed 22 December 2020].

Lamb, W.F. and Steinberger, J.K., 2017. Human well‐being and climate change mitigation. WIREs Climate Change, 8,
pp.e485. https://doi.org/10.1002/wcc.485.

Mansbridge, J., 2014. The role of state in governing the commons. Environmental Science & Policy, 36, pp.8–10.
https://doi.org/10.1016/j.envsci.2013.07.006.

Mcgillivray, M. and Clarke, M., eds., 2006. Understanding Human Well-being. Tokyo: UNU Press. p.386.

Millennium Ecosystem Assessment. 2005. Ecosystems and human well-being: synthesis/ current state and trends/ scenarios/
policy responses/ multiscale assessments. Washington, DC: Island Press.

Mitchell, T. and Maxwell, S., 2010. Defining climate compatible development. Climate & Development Knowledge
Network (CDKN) Policy Brief. Available at: https://cdkn.org/wp-content/uploads/2012/10/CDKN-CCD-
Planning_english.pdf [Accessed 03 September 2020].

Morone, P., Falcone, P.M. and Lopolito, A., 2019. How to promote a new and sustainable food consumption model: a fuzzy
cognitive map study. Journal of Cleaner Production, 208, pp.563–574. https://doi.org/10.1016/j.jclepro.2018.10.075.

Nápoles, G., Papageorgiou, E., Bello, R. and Vanhoof, K., 2016. On the convergence of sigmoid Fuzzy Cognitive Maps’.
Information Sciences, 349, pp.154–171. https://doi.org/10.1016/j.ins.2016.02.040.

Nerini, F.F., Tomei, J., To, L.S., Bisaga, I., Parikh, P., Black, M., Borrion, A., Spataru, C., Broto, V.C., Anandarajah, G. and
Milligan, B., 2018. Mapping synergies and trade-offs between energy and the Sustainable Development Goals. Nature
Energy, 3(1), pp.10–15. https://doi.org/10.1038/s41560-017-0036-5.

Nilsson, M., Griggs, D. and Visbeck, M., 2016. Policy: map the interactions between Sustainable Development Goals.
Nature, 534, pp.320–322. https://doi.org/10.1038/534320a.

Nunan, F., 2017. Making climate compatible development happen. London: Taylor & Francis. p.284.
eISBN:9781315621579.

O’Brien, K., 2018. Is the 1.5°C target possible? Exploring the three spheres of transformation. Current Opinion in
Environmental Sustainability, 31, pp.153–160. https://doi.org/10.1016/j.cosust.2018.04.010.

139
O’Neill, B.C., Oppenheimer, M., Warren, R., Hallegatte, S., Kopp, R.E., Pörtner, H.O., Scholes, R., Birkmann, J., Foden,
W., Licker, R., Mach, K.J., Marbaix, P., Mastrandrea, M.D., Price, J., Takahashi, K., van Ypersele, J.-P. and Yohe, G.,
2017. IPCC reasons for concern regarding climate change risks. Nature Climate Change, 7, pp.28–37.
https://doi.org/10.1038/nclimate3179.

O’Neill, D.W., Fanning, A.L., Lamb, W.F. and Steinberger, J.K., 2018. A good life for all within planetary boundaries.
Nature Sustainability, 1, pp.88–95. https://doi.org/10.1038/s41893-018-0021-4.

O’Neill, S., Williams, H.T.P., Kurz, T., Wiersma, B. and Boykoff, M., 2015. Dominant frames in legacy and social media
coverage of the IPCC Fifth Assessment Report. Nature Climate Change, 5, pp.380–385.
https://doi.org/10.1038/nclimate2535.

Omukuti, J., 2020. Challenging the obsession with local level institutions in country ownership of climate change adaptation.
Land Use Policy, 94, pp.104525. https://doi.org/10.1016/j.landusepol.2020.104525.

Ostrom, E., 2010. Polycentric systems for coping with collective action and global environmental change. Global
Environmental Change, 20, pp.550–557. https://doi.org/10.1016/j.gloenvcha.2010.07.004.

Özesmi, U. and Özesmi, L.S., 2004. Ecological models based on people’s knowledge: a multi-step fuzzy cognitive mapping
approach. Ecological Modelling, 176(1-2), pp.43–64. https://doi.org/10.1016/j.ecolmodel.2003.10.027.

Patterson, J., Schulz, K., Vervoort, J., van der Hel, S., Widerberg, O., Adler, C., Hurlbert, M. and Anderton, K., 2016.
politics of transformations towards sustainability. Environ. Innov. Soc. Transit.
http://dx.doi.org/10.1016/j.eist.2016.09.001.

Patterson, J.J., Thaler, T., Hoffmann, M., Hughes, S., Oels, A., Chu, E., Mert, A., Huitema, D., Burch, S. and Jordan, A.,
2018. Political feasibility of 1.5° C societal transformations: the role of social justice. Current Opinion in
Environmental Sustainability, 31, pp.1–9. https://doi.org/10.1016/j.cosust.2017.11.002.

Peake, S. and Ekins, P., 2016. Exploring the financial and investment implications of the Paris Agreement. Climate Policy,
17(1), pp.832–852. https://doi.org/10.1080/14693062.2016.1258633.

Pereira, I.P.C., Ferreira, F.A.F., Pereira, L.F., Govindan, K., Meidute-Kavaliauskiene, I. and Correia, R.J.C., 2020. A fuzzy
cognitive mapping-system dynamics approach to energy change impacts on the sustainability of small and medium-
sized enterprises. Journal of Cleaner Production, 256, pp.120154. https://doi.org/10.1016/j.jclepro.2020.120154.

Pörtner, H.-O., Karl, D.M., Boyd, P.W., Cheung, W.W.L., Lluch-Cota, S.E., Nojiri, Y., Schmidt, D.N. and Zavialov, P.O.,
2014. Ocean systems. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral
Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate
Change [Field, C.B., Barros, V.R., Dokken, D.J., Mach, K.J., Mastrandrea, M.D., Bilir, T.E., Chatterjee, M., Ebi, K.L.,
Estrada, Y.O., Genova, R.C., Girma, B., Kissel, E.S., Levy, A.N., MacCracken, S., Mastrandrea, P.R. and White, L.L.,
(eds.)]. Cambridge, United Kingdom and New York, NY, USA: Cambridge University Press. pp.411–484.

Reckien, D., 2014. Weather extremes and street life in India—implications of fuzzy cognitive mapping as a new tool for
semi-quantitative impact assessment and ranking of adaptation measures. Global Environmental Change, 26, pp.1–13.
https://doi.org/10.1016/j.gloenvcha.2014.03.005.

Rockstrom, J., Steffen, W., Noone, K., Persson, A., Chapin 3rd, F.S., Lambin, E.F., Lenton, T.M., Scheffer, M., Folke, C.,
Schellnhuber, H.J., Nykvist, B., de Wit, C.A., Hughes, T., van der Leeuw, S., Rodhe, H., Sörlin, S., Snyder, P.K.,
Costanza, R., Svedin, U., Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J., Walker, B., Liverman,
D., Richardson, K., Crutzen, P. and Foley, J.A., 2009. A safe operating space for humanity. Nature, 461, pp.472–475.
https://doi.org/10.1038/461472a.

Roy, J., Tschakert, P., Waisman, H., Abdul, H.S., Antwi-Agyei, P., Dasgupta, P., Hayward, B., Kanninen, M., Liverman, D.,
Okereke, C., Pinho, P.F., Riahi, K. and Suarez Rodriguez, A.G., 2018. Sustainable development, poverty eradication
and reducing inequalities. In: 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 [Masson-
Delmotte, V., Zhai, P., Pörtner, H.-O., Roberts, D., Skea, J., Shukla, P.R., Pirani, A., Moufouma-Okia, W., Péan, C.,
Pidcock, R., Connors, S., Matthews, J.B.R., Chen, Y., Zhou, X., Gomis, M.I., Lonnoy, E., Maycock, T., Tignor, M. and

140
Waterfield, T. (eds.)]. Cambridge, United Kingdom and New York, NY, USA: Cambridge University Press. pp.445–
538.

Ryan, R.M., Huta, V. and Deci, E.L. 2008. Living Well: A Self-Determination Theory Perspective on Eudaimonia. Journal
of Happiness Studies, 9, pp.139–170. https://doi.org/10.1007/s10902-006-9023-4.

Sachs, J.D., Schmidt-Traub, G., Mazzucato, M., Messner, D., Nakicenovic, N. and Rockström, J., 2019. Six Transformations
to achieve the Sustainable Development Goals. Nature Sustainability, 2, pp.805–814. https://doi.org/10.1038/s41893-
019-0352-9.

Santika, W.G., Anisuzzaman, M., Bahri, P.A., Shafiullah, G.M., Rupf, G.V. and Urmee, T., 2019. From goals to joules: A
quantitative approach of interlinkages between energy and the Sustainable Development Goals. Energy Research &
Social Science, 50, pp.201–214. https://doi.org/10.1016/j.erss.2018.11.016.

Schwartz, S.H., 1994. Are there universal aspects in the structure and contents of human values? Journal of Social Issues,
50, pp.19–45.

Scoones, I., Leach, M. and Newell, P., eds., 2015. The Politics of Green Transformations. London: Routledge. p.238.
eISBN:9781315747378.

Sellberg, M.M., Ryan, P., Borgström, S.T., Norström, A.V. and Peterson, G.D., 2018. From resilience thinking to Resilience
Planning: Lessons from practice. Journal of Environmental Management, 217, pp.906–918.
https://doi.org/10.1016/j.jenvman.2018.04.012.

Sen, A., 1993. Capability and Well-Being. In: Nussbaum, M. and Sen, A. The Quality of Life. Oxford: Clarendon Press.
ISBN-13:9780198287971.

Sen, A., 2009. The Idea of Justice. London: Penguin Books Ltd. p.496. eISBN:9780141969480.

Sen, A., 1985. Well-being, agency and freedom: The Dewey Lectures 1984. The Journal of Philosophy. 82, pp.169–221.
https://doi.org/10.2307/2026184.

Seneviratne, S.I., Rogelj, J., Séférian, R., Wartenburger, R., Allen, M.R., Cain, M., Millar, R.J., Ebi, K.L., Ellis, N., Hoegh-
Guldberg, O., Payne, A.J., Schleussner, C.-F., Tschakert, P. and Warren, R.F. 2018. The many possible climates from
the Paris Agreement’s aim of 1.5 °C warming. Nature, 558, pp.41–49. https://doi.org/10.1038/s41586-018-0181-4.

Shindell, D., Borgford-Parnell, N., Brauer, M., Haines, A., Kuylenstierna, J.C.I., Leonard, S.A., Ramanathan, V.,
Ravishankara, A., Amann, M. and Srivastava, L., 2017. A climate policy pathway for near-and long-term benefits.
Science, 356(6337), pp.493–494. https://doi.org/10.1126/science.aak9521.

Singh, P.K. and Chudasama, H., 2017. Assessing impacts and community preparedness to cyclones: a fuzzy cognitive
mapping approach. Climatic Change, 143, pp.337–354. https://doi.org/10.1007/s10584-017-2007-z.

Singh, P.K. and Chudasama, H., 2020. Evaluating Poverty Alleviation Strategies in a Developing Country. PLoS ONE,
15(1), pp.e0227176. https://doi.org/10.1371/journal.pone.0227176.

Singh, P.K. and Chudasama, H., 2021. Pathways for climate change adaptations in arid and semi-arid regions. Journal of
Cleaner Production, 284:124744. https://doi.org/10.1016/j.jclepro.2020.124744.

Singh, P.K. and Nair, A., 2014. Livelihood vulnerability assessment to climate variability and change using fuzzy cognitive
mapping approach. Climatic Change, 127, pp.475–491. https://doi.org/10.1007/s10584-014-1275-0.

Sirgy, M.J., 2012. The psychology of quality of life: Hedonic well-being, life satisfaction, and eudaimonia. Second.
Netherlands: Springer. p.622. eISBN:978-94-007-4405-9.

Solecki, W., Rosenzweig, C., Dhakal, S., Roberts, D., Barau, A.S., Schultz, S. and Ürge-Vorsatz, D., 2018. City
transformations in a 1.5 C warmer world. Nature Climate Change, 8(3), pp.177–181. https://doi.org/10.1038/s41558-
018-0101-5.

Sørensen, E., 2017. Political innovations: innovations in political institutions, processes and outputs. Public Management
Review, 19(1), pp.1–19, https://doi.org/10.1080/14719037.2016.1200661.

141
Sørensen, E. and Torfing, J. 2018. Governance on a bumpy road from enfant terrible to mature paradigm. Critical Policy
Studies 12(3), pp.350–359. https://doi.org/10.1080/19460171.2018.1437461.

Spaiser, V., Ranganathan, S., Swain, R.B. and Sumpter, D.J.T., 2017. The sustainable development oxymoron: quantifying
and modelling the incompatibility of sustainable development goals. International Journal of Sustainable Development
& World Ecology, 24(6), pp.457–470. https://doi.org/10.1080/13504509.2016.1235624.

Steffen, W., Richardson, K., Rockström, J., Cornell, S.E., Fetzer, I., Bennett, E.M., Biggs, R., Carpenter, S.R., de Vries, W.,
de Wit, C.A., Folke, C., Gerten, D., Heinke, J., Mace, G.M., Persson, L.M., Ramanathan, V., Reyers, B. and Sörlin, S.,
2015. Planetary boundaries: Guiding human development on a changing planet. Science, 347(6223), pp.1259855.
https://doi.org/10.1126/science.1259855.

Stiglitz J., Sen A. and Fitoussi J.-P., 2009. Report by the commission on the measurement of economic performance and
social progress. Paris: Commission on the Measurement of Economic Performance and Social Progress. p.291. [online]
Available at: https://www.cpc.unc.edu/projects/rlms-hse/publications/1921 [Accessed 14 July 2020].

Stylios, C.D., Georgopoulos, V.C. and Groumpos, P.P., 1997. The use of fuzzy cognitive maps in modeling systems. In:
Proceedings of 5th IEEE Mediterranean conference on control and systems, Paphos, Cyprus, pp.21–23.

UN ESCAP, 2019. Asia and the Pacific SDG Progress Report. United Nations (UN) Economic and Social Commission for
Asia and the Pacific (ESCAP). Bangkok: United Nations Publications. p.73. eISBN:978-92-1-004173-7. [online]
Available at:
https://www.unescap.org/sites/default/files/publications/ESCAP_Asia_and_the_Pacific_SDG_Progress_Report_2019.p
df [Accessed 19 August 2020].

UN, 2015. Transforming Our World: The 2030 Agenda for Sustainable Development. United Nations (UN). [online]
Available at:
https://sustainabledevelopment.un.org/content/documents/21252030%20Agenda%20for%20Sustainable%20Developm
ent%20web.pdf [Accessed 7 August 2020].

UNEP, 2016: The Adaptation Finance Gap Report 2016. Nairobi, Kenya: United Nations Environment Programme (UNEP).
[online] Available at: https://climateanalytics.org/media/agr2016.pdf [Accessed 20 October 2020].

UNFCCC, 2015: The Paris Agreement. United Nations Framework Convention on Climate Change (UNFCCC), p.27.
[online] Available at: https://unfccc.int/sites/default/files/english_paris_agreement.pdf [Accessed 18 September 2020].

Wolf, J., Allice, I. and Bell, T., 2013. Values, climate change, and implications for adaptation: Evidence from two
communities in Labrador, Canada. Global Environmental Change, 23(2), pp.548–562.
https://doi.org/10.1016/j.gloenvcha.2012.11.007.

Ziv, G., Watson, E., Young, D., Howard, D.C., Larcom, S.T. and Tanentzap, A.J., 2018. The potential impact of Brexit on
the energy, water and food nexus in the UK: A fuzzy cognitive mapping approach. Applied Energy, 210, pp.487–498.
https://doi.org/10.1016/j.apenergy.2017.08.033.

142

You might also like