Eu NBS

Download as pdf or txt
Download as pdf or txt
You are on page 1of 373

EVALUATING THE IMPACT OF

NATURE-BASED
SOLUTIONS
Independent
A Handbook for Practitioners
Expert
Report

Knowledge building Place


for sustainable urban regeneration
Green space Health and
transformation
management well-being

Water
Participatory planning Climate resilience management New economic
and governance opportunities and
Biodiversity
green jobs
enhancement

Natural and Social justice and


climate hazards social cohesion

Air quality

Research and
Innovation
Evaluating the Impact of Nature-based Solutions: A Handbook for Practitioners
European Commission
Directorate-General for Research and Innovation
Directorate C — Healthy Planet
Unit C3 — Climate and Planetary Boundaries

Contact [email protected]
[email protected]
Email [email protected]
[email protected]
European Commission
B-1049 Brussels

Manuscript completed in March 2021.


First edition.

This document has been prepared for the European Commission, however it reflects the views only of the authors, and the European Commission
is not liable for any consequence stemming from the reuse of this publication.
More information on the European Union is available on the internet (http://europa.eu).

PDF ISBN 978-92-76-22821-9 doi:10.2777/244577 KI-04-20-586-EN-N

Luxembourg: Publications Office of the European Union, 2021


© European Union, 2021

The reuse policy of European Commission documents is implemented based on Commission Decision 2011/833/EU of 12
December 2011 on the reuse of Commission documents (OJ L 330, 14.12.2011, p. 39). Except otherwise noted, the reuse of this
document is authorised under a Creative Commons Attribution 4.0 International (CC-BY 4.0) licence (https://creativecommons.
org/licenses/by/4.0/). This means that reuse is allowed provided appropriate credit is given and any changes are indicated.
For any use or reproduction of elements that are not owned by the European Union, permission may need to be sought directly from the respective
rightholders.
Image credits:
cover: © MicroOne # 305386384, 2019. Source: stock.adobe.com
EUROPEAN COMMISSION

EVALUATING THE IMPACT OF

NATURE-BASED
SOLUTIONS
A Handbook for Practitioners

Adina Dumitru and Laura Wendling, Eds.

Directorate-General for Research and Innovation


2021 Healthy Planet - Climate and Planetary Boundaries
Table of Contents
FOREWORD .................................................................................................. 6
LIST OF ABBREVIATIONS .......................................................................... 10
1. INTRODUCTION ..................................................................................... 16
1.1 What are Nature-based Solutions? ......................................................... 17
1.2 NBS in European and International policy frameworks .............................. 20
1.2.1 NBS in the European policy context ................................................. 20
1.2.2 NBS in an International policy context .............................................. 23
1.3 Purpose of the NBS Impact Evaluation Handbook ..................................... 25
1.3.1 Handbook aim ............................................................................... 25
1.3.2 Intended audience of this handbook................................................. 26
1.3.3 How this handbook was developed................................................... 27
1.4 Content of this handbook ...................................................................... 34
1.5 Conclusions ......................................................................................... 37
1.6 References .......................................................................................... 38

PROFILE: NATURE4CITIES ............................................................................. 40


PROFILE: NATURVATION ............................................................................... 42
PROFILE:THINK NATURE................................................................................ 44

2. PRINCIPLES GUIDING NBS PERFORMANCE AND IMPACT EVALUATION . 46


2.1 Introduction and definitions................................................................... 47
2.1.1 The concept of effectiveness ........................................................... 51
2.2 Decision-making context and impact evaluations: from needs to indicators . 53
2.3 Principles for the development of impact monitoring and evaluation plans ... 57
2.3.1 Steps ........................................................................................... 57
2.3.2 Principles ...................................................................................... 58
2.4 Capitalising on existing experiences and remaining critical concerns ........... 64
2.4.1 Challenges and gaps in current monitoring and evaluation efforts ........ 64
2.4.2 Key messages from existing projects ............................................... 67
2.5 References .......................................................................................... 68
PROFILE: CONNECTING NATURE .................................................................... 70
PROFILE: GROW GREEN ................................................................................ 72
PROFILE: UNALAB ......................................................................................... 74
PROFILE: URBAN GREENUP ............................................................................ 76

3. APPROACHES TO MONITORING AND EVALUATION STRATEGY


DEVELOPMENT ........................................................................................... 78
3.1 Introduction: developing robust impact assessment plans ......................... 79

2
3.2 A step by step approach to developing robust monitoring and evaluation plans
for NBS .................................................................................................... 80
3.3 Robust assessment and co-production: a necessary relationship ................ 90
3.4 Innovative tools for monitoring and evaluation of nature-based solutions.... 96
3.4.1 Reflexive monitoring – Connecting Nature project .............................. 96
3.4.2 iAPT (Impact Assessment Planning Tool) – Connecting Nature project .. 99
3.4.3 Urban GreenUP Tool – Urban GreenUP project .................................. 100
3.5 Conclusions ........................................................................................ 102
3.6 References ......................................................................................... 103
PROFILE: CLEVER CITIES ............................................................................. 106
PROFILE: PROGIREG .................................................................................... 108
PROFILE: EDICITNET.................................................................................... 110
PROFILE: URBINAT ...................................................................................... 112

4. INDICATORS OF NBS PERFORMANCE AND IMPACT .............................. 114


4.1 Societal challenge areas addressed by NBS ............................................ 116
4.2 Recommended and Additional indicators for NBS impact assessment ......... 120
4.2.1 Climate Resilience ........................................................................ 124
4.2.2 Water Management ...................................................................... 128
4.2.3 Natural and Climate Hazards .......................................................... 132
4.2.4 Green Space Management ............................................................. 137
4.2.5 Biodiversity Enhancement.............................................................. 142
4.2.6 Air Quality ................................................................................... 145
4.2.7 Place Regeneration ....................................................................... 148
4.2.8 Knowledge and Social Capacity Building for Sustainable Urban
Transformation ..................................................................................... 151
4.2.9 Participatory Planning and Governance ............................................ 153
4.2.10 Social Justice and Social Cohesion ................................................ 156
4.2.11 Health and Wellbeing .................................................................. 158
4.2.12 New Economic Opportunities and Green Jobs ................................. 163
4.3 Conclusions ........................................................................................ 168
4.3.1 Summary of the indicator framework presented ............................... 168
4.3.2 Emerging concerns and further development needs .......................... 168
4.4 References ......................................................................................... 169
PROFILE: CLEARING HOUSE.......................................................................... 173
PROFILE: REGREEN...................................................................................... 175

5. APPLICATION OF THE NBS IMPACT EVALUATION FRAMEWORK: NBS


PERFORMANCE AND IMPACT EVALUATION CASE STUDIES ...................... 177
5.1 Introduction to holistic NBS impact assessment using the framework of
recommended indicators ........................................................................... 179
5.1.1 Recommended indicators case study from Tampere, Finland .............. 181

3
5.1.2 Recommended indicators case study from Valladolid, Spain ............... 183
5.1.3 Recommended indicators case study from Guildford, UK .................... 186
5.1.4 Recommended indicators case study from Genk, Belgium .................. 189
5.2 Case studies illustrating the ‘story of an indicator’ for some of the additional
indicators ................................................................................................ 196
5.2.1 Climate Resilience – Urban heat Island incidence .............................. 196
5.2.2 Natural and climate hazards – Flood risk ......................................... 199
5.2.3 Green space management – Walkability .......................................... 203
5.2.4 Green space management – Annual trend in vegetation cover............ 205
5.2.5 Green space management – ESTIMAP nature-based recreation .......... 211
5.2.6 Green space management – Land composition ................................. 215
5.2.7 Biodiversity Enhancement – Number of conservation priority species .. 219
5.2.8 Air Quality – Trends in NOx and SOx emissions ................................ 221
5.2.9 Knowledge and Social Capacity Building for Sustainable Urban
Transformation – Connectedness to nature .............................................. 223
5.2.10 Social Justice and Social Cohesion – Perceived social support ........... 224
5.2.11 Health and Wellbeing – Prevalence, incidence, and morbidity of chronic
stress .................................................................................................. 226
5.2.12 Health and Wellbeing – Perceived chronic loneliness........................ 229
5.3 Conclusions ........................................................................................ 232
PROFILE: NAIAD .......................................................................................... 233
PROFILE: OPERANDUM ................................................................................. 235
PROFILE: PHUSICOS .................................................................................... 237
PROFILE: RECONECT .................................................................................... 239

6. NBS FOR DISASTER RISK REDUCTION ................................................. 241


6.1 NBS and Disaster Risk Reduction .......................................................... 242
6.2 Basics of risk analysis, risk reduction measures, resilience and effectiveness
.............................................................................................................. 243
6.3 Indicators and methodologies for measuring NBS effectiveness indicators in
DRR context ............................................................................................ 248
6.4 Case study #1 – from indicators assessment to integration and decision-aiding
for flood risk management ......................................................................... 254
6.4.1 Context and global framework for assessment of NBS effectiveness .... 254
6.4.2 Indicators for assessment of technical, physical and economic efficacy of
flood mitigation strategies including NBS ................................................. 255
6.5 Case study #2 – a green barrier to reduce the risk of floods due to snowmelt
and extreme rainfall ................................................................................. 258
6.5.1 General background and hazard type .............................................. 258
6.5.2 Co-benefits of the proposed NBS .................................................... 259
6.5.3 Indicators for the NBS performance assessment ............................... 260
6.6 Case study #3 – landslides and debris flows........................................... 264
6.7 Case study #4 – floods in dense urban environments .............................. 268

4
6.8 Concluding remarks ............................................................................ 270
6.9 References ......................................................................................... 270
PROFILE: MAES ........................................................................................... 273
PROFILE: ENROUTE ..................................................................................... 275

7. DATA REQUIREMENTS .......................................................................... 277


7.1 Data terminology, definitions and key concepts ...................................... 279
7.1.1 Spatial versus non-spatial data ...................................................... 279
7.1.2 Baseline data ............................................................................... 280
7.1.3 Control data................................................................................. 280
7.1.4 Acquisition regime ........................................................................ 281
7.1.5 Spatial scale of analysis................................................................. 282
7.1.6 Processing level............................................................................ 283
7.1.7 Data generation and collection methods .......................................... 284
7.2 Environmental data of relevance for NBS monitoring and assessment ........ 287
7.2.1 Remote sensing (RS) and Earth Observation (EO) ............................ 288
7.2.2 In-situ observations and ground measurements ............................... 296
7.2.3 Surveys ....................................................................................... 300
7.3 Socio-economic, demographic and behavioural datasets for NBS monitoring and
assessment: Methods and sources .............................................................. 301
7.3.1 Quantitative, qualitative and map-based surveys .............................. 303
7.3.2 Population observations................................................................. 306
7.4 Data sources for the assessment of changes to health and wellbeing ......... 307
7.5 Predicting the present and future impacts of NBS with modelling techniques
.............................................................................................................. 311
7.6 Mimicking the impacts of NBS: how laboratory data can help ................... 326
7.7 Engaging the community in the data collection process: citizen science and its
role in NBS monitoring .............................................................................. 326
7.8 Data integration ................................................................................. 329
7.9 Baseline assessment ........................................................................... 339
7.10 Data adequacy and related aspects ..................................................... 344
7.10.1 Data gaps and irregularities ......................................................... 348
7.10.2 Data granularity and resolution .................................................... 349
7.10.3 Data accuracy ............................................................................ 351
7.10.4 Biases, main error sources, and data reliability ............................... 353
7.10.5 Data accessibility ........................................................................ 353
7.10.6 Metadata and data standardization ............................................... 355
7.11 Conclusion ........................................................................................... 357
7.12 References .......................................................................................... 361

5
FOREWORD

Urban expansion and densification brings both opportunities and challenges.


Regeneration of urban areas is therefore a significant priority, which needs to
take into account environmental quality, social justice and sustainable
development. Transforming cities and regions into vibrant, sustainable and
resilient living places has become a key global priority. This is reflected in
numerous policy initiatives at local, region al and national scale, and
internationally through the UN Sustainable Development Goals (particularly SDG
11). Together these are part of a global call to rethink and redesign urban
environments through innovative solutions that address multiple issues.

The EU Research and Innovation policy agenda on Nature-based Solutions and


Re-naturing Cities defines nature-based solutions to societal challenges as
“solutions that are inspired and supported by nature, which are cost-effective,
simultaneously provide environmental, social and economic benefits and help
build resilience. Such solutions bring more, and more diverse, nature and natural
features and processes into cities, landscapes and seascapes, through locally
adapted, resource-efficient and systemic interventions” 1. Nature-based solutions
(NBS) intrinsically provide biodiversity benefits and support the delivery of
ecosystem services; however, there is increasing recognition of the multitude of
environmental, social and economic co-benefits delivered by NBS.

The objective of this handbook is to support the adoption of common indicators


and methods for assessing the performance and impact of diverse types of NBS.
The handbook is designed to be relevant for NBS implemented across a wide
geographic area and at a multitude of scales. The integrated NBS assessment
framework presented in the handbook has been developed with the three-fold
objective of:
• Serving as a reference for relevant EU policies and activities;

• Orienting urban practitioners in developing robust impact evaluation


frameworks for nature-based solutions at different scales; and,

• Providing a comprehensive set of indicators and methodologies.

This handbook is intended to serve as a guide to the development and


implementation of scientifically-valid monitoring and evaluation plans for the
evaluation of NBS impacts (Figure 1). We begin by defining NBS in the context of
global challenges and key policy instruments (Chapter 1). Subsequent chapters
guide the reader through the development and execution of robust NBS
monitoring and evaluation plans (Chapter 2 and Chapter 3), the selection
(Chapter 4 and Appendix of Methods) and application (Chapter 5) of impact
indicators, the use of NBS in Disaster Risk Reduction (DRR; Chapter 6), and the
acquisition and management of relevant data (Chapter 7).

1
https://ec.europa.eu/info/research-and-innovation/research-area/environment/nature-based-solutions_en

6
Why do we need a coordinated approach to NBS impact monitoring? Chapter 1
describes how the development of robust monitoring and evaluation frameworks
to assess NBS impacts enables cities and regions to assess the strengths and
weaknesses of specific interventions in achieving strategic goals, understand the
realised benefits and trade-offs, and sustainably manage NBS in the long term.
Chapter 1 also describes how monitoring and evaluation can help to build the
case for investments in NBS.

How do monitoring and evaluation contribute to evidence-based policy-making


and policy learning? Monitoring and evaluation tells us whether an NBS functions
as desired by providing evidence of its ability to achieve specific outcomes.
Chapter 2 describes the principles that guide NBS performance and impact
evaluation to support the development of an appropriate, scientifically robust NBS
monitoring and evaluation plan. The chapter presents general steps along with
advice on how these steps can be tailored to suit a specific NBS context.

Figure1. Overall structure and content of this handbook.

7
Chapter 3 further elaborates the steps in the development of monitoring and
evaluation plans. The development of local NBS monitoring and evaluation
strategies are illustrated by a series of case studies from several EU H2020
projects. In particular, Chapter 3 emphasises the connection between NBS
evaluation and monitoring plans and the processes of knowledge co-production
and NBS co-management.

How is impact measured? The impacts of NBS can be assessed quantitatively


and/or qualitatively by adopting indicators, a set of variables providing the means
to assess particular attributes to meet an explicit objective. Identification and
selection of specific indicators to evaluate NBS can seem a daunting prospect due
the vast selection of potential indicators and their specific metrics. The buffet-
style overview of indicators in this handbook helps the reader select the
appropriate indicators. The handbook builds upon and expands the EKLIPSE
Expert Working Group Impact evaluation framework. Chapter 4 presents a suite
of Recommended and Additional indicators to evaluate NBS impact across the
following 12 societal challenge areas:
1. Climate Resilience
2. Water Management
3. Natural and Climate Hazards
4. Green Space Management
5. Biodiversity
6. Air Quality
7. Place Regeneration
8. Knowledge and Social Capacity Building for Sustainable Urban Transformation
9. Participatory Planning and Governance
10. Social Justice and Social Cohesion
11. Health and Well-being
12. New Economic Opportunities and Green Jobs

In addition to the identification and classification of NBS impact indicators across


each of the 12 identified societal challenge areas, a range of methodological
approaches are presented in the accompanying Evaluating the Impact of Nature-
based Solutions: Appendix of Methods. The Appendix of Methods provides a
brief description of each indicator determination method, along with guidance for
end-users about the appropriateness, advantages and drawbacks of each method
in different contexts.

How does it all fit together? Chapter 5 presents a number of different case
studies to further illustrate the selection and application of indicators for impact
evaluation of different types of NBS implemented across a range of scales and in
diverse environments. The examples display how indicators can be used together
to address specific issues with the aim to inspire other cities and regions in
developing robust monitoring and evaluation frameworks and facilitate evidence-
based urban policy-making for NBS.

Chapter 6 details the use of NBS in ecosystem-based disaster risk reduction


(Eco-DRR) schemes, outlining the components of risk and the potential impacts

8
of NBS on risks due to natural phenomena. The use of NBS for DRR is illustrated
by a series of case studies focused on large-scale hydro-meteorological risk
reduction.

Chapter 7 provides an overview of data types, sources and techniques for the
generation of data to monitor and assess the impacts of NBS. An understanding
of different types of data, their sources and use is core to the development of
robust monitoring and evaluation plans.

The handbook supports practitioners to independently design and implement NBS


impact evaluation schemes. The indicators and methods of NBS impact
assessment presented reflect the state of the art in scientific research on impacts
of nature-based solutions and are valid and standardised methods of assessment.
The selection is not exhaustive, but acts as a European reference framework on
NBS impact evaluation and monitoring. The handbook synthesises information
concerning the current state of play in the implementation of evaluation
frameworks, as fostered by the European agenda on climate change adaptation
and disaster risk reduction, including the re-naturing of cities and urban
transformation towards sustainable, liveable, healthy and just cities.

This handbook was collaboratively developed by the NBS Impact Evaluation


Taskforce, a clustering initiative by the EU Commission to capitalise on synergies
between H2020 funded projects relating to NBS. The handbook expands on the
pioneering work of the EKLIPSE Working Group on Nature-based Solutions to
Promote Climate Resilience in Urban Areas.

These Horizon2020 funded projects and collaborating institutions contributed to


the NBS Impact Evaluation Taskforce that prepared this handbook (in
alphabetical order): CLEARING HOUSE; CLEVER Cities; CONNECTING Nature;
EdiCitNet; EEA; GROW GREEN; JRC; MAES/EnRoute; NAIAD; Nature4Cities;
Naturvation; OPERANDUM; PHUSICOS; proGIreg; RECONECT; REGREEN; Think
Nature; UNaLab; URBAN GreenUP; and, URBiNAT. The taskforce has relied on the
input of more than 150 European researchers and over 60 European cities and
regions involved in these projects. We thank all authors, lead authors and
coordinating lead authors for their hard work and commitment to developing the
handbook, and the European Commission for their support throughout the
development of this work.

We hope that this handbook is helpful to those who make the difference in the
field - practitioners, planners and decision-makers who implement NBS. Let this
handbook inspire your work.

Rik De Vrees
Adina Dumitru
Sebastian Eiter
Laurence Jones
Laura Wendling
Marianne Zandersen

9
LIST OF ABBREVIATIONS

ABM Agent-based model


ADCIRC Advanced circulation model
ADHD Attention deficit hyperactivity disorder
ANK Atlas of Natural Capital
API Application programming interface
AQP Air quality pollutant
ARIES Artificial Intelligence for Ecosystem Services
ART Attention Restoration Theory
AVHRR Advanced very high-resolution radiometer
B£ST Benefits Estimation Tool
BC Black carbon
BI Blue infrastructure
BGI Blue-green infrastructure
BISE Biodiversity Information System for Europe
BMI Body mass index
BMPs Best management practices
bVOC Biogenic volatile organic compound
CA Cellular automata
CBA Cost-benefit analysis
CCA Climate change adaptation
CH4 Methane
CIF Common Implementation Framework
CNS Connectedness to nature scale
CO Carbon monoxide
CO2 Carbon dioxide
CO2e Carbon dioxide equivalent
CORDEX Coordinated Regional Climate Downscaling Experiment
CVD Cardiovascular disease
DEM Discrete element method
DRMKC Disaster Risk Management Knowledge Centre
DRR Disaster risk reduction
EbA Ecosystem-based adaptation

10
Eco-DRR Ecosystem-based disaster risk reduction
EC European Commission
ECMWF European Centre for Medium-Range Weather Forecasts
ECS Edible City Solutions
ECV Essential climate variable
EE Ecological engineering
EEA European Environment Agency
EO Earth observation
ERA40 Re-analysis of meteorological data from September
1957 to August 2002 produced by ECMWF
ESA European Space Agency
ESM European Settlement Map
ESS Ecosystem services
ESTIMAP Ecosystem Services Mapping tool
EU European Union
Eurostat Statistical Office of the European Union
FAIR Findability, accessibility, interoperability and reusability
of data
FEV Flood excess volume
FRAME Fine Resolution Atmospheric Multi-species Exchange
model
FRC Front-runner city
FUA Functional urban area
GCM General circulation model
GDPR General Data Protection Regulation
GEE Google Earth engine
GHG Greenhouse gas
GHSL Global Human Settlement Layer
GI Green infrastructure
GIS Geographic Information System
GLEON Global Lake Ecological Observatory Network
GVA Gross value added
H2020 Horizon 2020 framework programme
HEC Hydrologic Engineering Center
HEE Hydrological extreme event

11
HFA Hyogo Framework for Action
HMR Hydro-meteorological risk
IACS Integrated Agriculture and Control System
ICOS Integrated Carbon Observation System
ILO International Labour Organization
INSPIRE Infrastructure for Spatial Information in Europe
InVEST Integrated Valuation of Ecosystem Services and
Tradeoffs
IPAQ International physical activity questionnaire
IPCC Intergovernmental Panel on Climate Change
ISO International Organization for Standardization
IUCN International Union for the Conservation of Nature
IVR Immersive virtual reality
JRC Joint Research Centre
KIP INCA Integrated system of Natural Capital and Ecosystem
Services accounting
KPI Key performance indicator
LAI Leaf area index
LID Low-impact development
LiDAR Light detection and ranging
LL Living Lab
LM Landscape mosaic
LUCI Land Utilisation Capability Indicator
LUE Land Use Efficiency
LUISA Land Use-based Integrated Sustainability Assessment
LULC Land use and land cover
LUT Look-up tables
M&E Monitoring and evaluation
MAES Mapping and Assessment on Ecosystems and their
Services
MCDA Multicriteria decision analysis
MODIS Moderate resolution imaging spectroradiometer
NBS Nature-based solution
NC Natural capital
NDVI Normalised Difference Vegetation Index

12
NGO Non-governmental organisation
NO2 Nitrogen dioxide
NO3-N Nitrate-nitrogen
NOx Nitrogen oxides
NUTS Nomenclature of Territorial Units for Statistics
NWRM Natural Water Retention Measures
O3 Ozone
OAL Open Air Laboratory
OECD Organisation for Economic Cooperation and
Development
OGC Open Geospatial Consortium
OS Opportunity spectrum
OSGeo Open Source Geospatial Foundation
OSM Open Street Map
PAH Polycyclic aromatic hydrocarbon
PLS Partial least square
PM Particulate matter
PM2.5 Particulate matter <2.5 µm in diameter
PM10 Particulate matter <10 µm in diameter
PPGIS Public participation geographic information system
PPP Public-private partnership
ROI Return on investment
RP Recreation potential
ROS Recreation Opportunity Spectrum
RS Remote sensing
RUP Re-naturing Urban Plan
SAR Synthetic aperture radar
SCI Site of community importance
SD System dynamics
SDG Sustainable Development Goal
SEA Strategic environmental assessment
SEDAC Socioeconomic Data and Applications Centre
SES Social-ecological systems
SFDRR Sendai Framework for Disaster Risk Reduction

13
SMART Specific, Measurable, Attributable, Realistic, Targeted
SO2 Sulphur dioxide
SOx Sulphur oxides
SolVES Social Values for Ecosystem Services
SOPARC System for Observing Play and Recreation in
Communities
SPA Special protection area
SRA Strategic Research Agenda
SROI Social return on investment
SRT Stress Recovery Theory
SuDs Sustainable urban drainage systems
SWAN Simulative Waves Nearshore model
SWAT Soil Water Assessment Tool
SWMM Storm Water Management Model
TC Technical Committee
TEEB The Economics of Ecosystems and Biodiversity
TESSA Toolkit for Ecosystem Service Site-based Assessment
TF Taskforce
TOPHEE Approach combining indicators for technical, physical,
organizational, environmental, social/human and
economic features
TSS Total suspended solids
UCDB Urban Centres Database
UCM Urban canopy model
UCS Urban Carbon Sink
UF Urban forestry
UGI Urban green infrastructure
UHI Urban Heat Island
ULL Urban Living Lab
UN United Nations
UNA Urban Nature Atlas
UNEP United Nations Environment Programme
UNISDR United Nations International Strategy for Disaster
Reduction
UTCI Universal Thermal Comfort Index

14
VGI Volunteered geographic information
VOC Volatile organic compound
WCDRR World Conference on Disaster Risk Reduction
WEAP Water Evaluation and Planning model
WHO World Health Organisation
WMO World Meteorological Organization
WSN Wireless sensor network
WSUD Water-sensitive urban design
WRF Weather Research and Forecasting Model
YoLL Years of life lost

15
1 INTRODUCTION
Coordinating Lead author
Sgrigna, G.

Lead authors
Sgrigna, G., López-Gunn, E., Dubovik, M.

Contributing authors
Di Sabatino, S., Kumar, P., Feliu, E., Ruangpan, L., San Jose, E., Sanchez, R., Van
Cauwenbergh, N., Vojinovic, Z., Wendling, L.

Summary

What is this chapter about?

This chapter introduces the aim of the NBS Impact Evaluation Handbook as a
reference for evaluating the impacts of nature-based solutions (NBS). It provides
a general framework on the value of NBS to the community, investors, and policy
makers, and illustrates how the NBS impact evaluation framework can be used.
Chapter 1 describes the global context in which NBS operate. Two infographics
help visualise the definition of NBS and provide an in-depth explanation of the
concept’s origin and evolution. Another infographic describes the full life cycle of
NBS including monitoring, evaluation, and cost-benefit analysis. The chapter
concludes by describing the content of each section of the handbook.

16
Chapter 1 illustrates how an impact evaluation framework supports:

1) Policy evaluation and the achievement of policy and regulatory goals;


2) Social accountability, so that citizens’ concerns are taken into account;
and,
3) Investment in NBS, including the comparison of NBS impacts with those
of other technical engineered approaches.

How can I use this chapter in my work with NBS?

Chapter 1 provides fundamental background information on the concept of NBS,


its adoption and the benefits of assessing NBS design, uptake, and
implementation.

When can I use this knowledge in my work with NBS?

It is particularly useful during the early stages, to understand the framing, when
you start planning NBS implementation and the monitoring and evaluation
framework.

How does this chapter link with the other parts of the handbook?

This chapter frames the content of the NBS Impact Evaluation Handbook and
provides an overall guide to its different sections.

1.1 What are Nature-based Solutions?

The concept of nature-based solutions embodies new ways to approach socio-


ecological adaptation and resilience, with equal reliance upon social,
environmental and economic domains. Nature-based solutions (NBS) were clearly
described for the first time in the final report of the Horizon 2020 Expert Group
(EC, 2015). The European Commission defines NBS as solutions that are “inspired
and supported by nature, which are cost-effective, simultaneously provide
environmental, social and economic benefits and help build resilience. Such
solutions bring more, and more diverse, nature and natural features and
processes into cities, landscapes and seascapes, through locally adapted,
resource-efficient and systemic interventions.” 2 Inherent in this definition is the
idea that NBS must benefit biodiversity and support the delivery of a range of
ecosystem services. Similarly, the International Union for Conservation of Nature
(IUCN) defines NBS as “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” 3.

2
https://ec.europa.eu/info/research-and-innovation/research-area/environment/nature-based-solutions_en
3
https://www.iucn.org/commissions/commission-ecosystem-management/our-work/nature-based-solutions

17
The NBS concept, as reported by Escobedo et al. (2019), is the evolution of terms
used previously to express similar ideas: urban forestry (UF); green and blue
infrastructure (GI, BI); and ecosystem services (ESS). Eisenberg et al. (2018)
and Ruangpan et al. (2020) identify additional concepts and practices that can
be broadly placed under the umbrella of NBS: ecosystem-based adaptation
(EbA), ecosystem-based disaster risk reduction (Eco-DRR), blue–green
infrastructure (BGI), low-impact development (LID), best management practices
(BMPs), water-sensitive urban design (WSUD), sustainable urban drainage
systems (SuDs), and ecological engineering (EE). With respect to NBS, these
existing concepts are applicable across strategic, spatial planning, soft
engineering, and performance dimensions (Figure 1-1).

Experts with different backgrounds view NBS through various disciplinary lenses.
Dorst et al. (2019) describe NBS as “interventions based on nature that are
envisaged to address sustainability challenges such as resource shortages, flood
and heat risks and ecosystem degradation caused by processes of urbanization
and climate change”. Kabisch et al. (2016) underline the connection of NBS with
“the maintenance, enhancement, and restoration of biodiversity and ecosystems
as a means to address multiple concerns simultaneously”. In contrast,
Frantzeskaki et al. (2017) view NBS in a social-ecological context, noting that
“transition initiatives as actor configurations that establish, experiment and
localise nature-based solutions shift them from ‘solutions’ to social
configurations, making nature-based solutions the new ‘urban commons of
sustainability’...”. A recent editorial about NBS within the Nature journal stated
that “the concept it represents is of vital and urgent significance. As the grand
challenges that face society continue to build, so does the need for
multidisciplinary, evidence-based strategies to, for example, protect water
supplies, address habitat loss and mitigate and adapt to climate change” ('Natural
language: the latest attempt to brand green practices is better than it sounds',
2017). In short, NBS provide integrated, multifunctional solutions to many of our
current urban and rural challenges through the use of nature and natural
processes.

18
Figure 1-1. Nature-based solutions as an umbrella concept and the relation of NBS to key existing
concepts. EbA = ecosystem based adaptation; Eco-DRR = ecosystem-based disaster risk reduction; GI =
green infrastructure; BI = clue infrastructure; GBI = green-blue infrastructure; UF = urban forestry; SuDS
= sustainable urban drainage systems; EE = ecological engineering; BMPs = best management practices;
LID = low-impact design; WSUD = water-sensitive urban design; ESS = ecosystem services.

The application of NBS is the deliberate inclusion of natural system processes


within human environments to obtain relevant outcomes in the form of ecosystem
services. For example, a well-managed forest can provide multiple ESS, including
provisioning, regulating and cultural ecosystem services. Provisioning services
provided by a forest may include timber, fuel, fibre, and/or food. Climate
regulation is one example of a regulating service provided by forests due to
evapotranspiration and shading of land surfaces (cooling), and the removal and
fixation of atmospheric CO2 within tree biomass. The leaves and roots of trees
can also intercept and lessen rainfall runoff and reduce the impact of flooding
(flood regulation), creating a natural buffer. The cultural services provided by
natural areas are increasingly recognised, including benefits such as enhanced
mental well-being, increased recreational value, educational opportunities, new
job niches in nature-based enterprises or less tangible aesthetic and spiritual
benefits.

19
NBS are characterised by their capacity to simultaneously address several
societal challenges in terms of primary benefits and co-benefits, or ecosystem
services. Among other positive impacts, such as enhanced resilience to the
impacts of climate change or increased biodiversity, one of the common
denominators of NBS is the concept of sustainability. The implementation of NBS
in human environments could be considered as a fundamental tool capable of
sustaining human life and activities over time in a way that is compatible with
the planetary boundaries (Rockström et al., 2009); a “green – blue pedal” in the
hands of policy makers, administrators and practitioners. In other words, NBS
provide opportunity to enhance and maintain the liveability of human settlements
for current and future generations.

The development of new urban environments based on sustainability are


undermined by the standard models of urbanisation processes (Lafortezza and
Sanesi 2019). There is growing evidence regarding the benefits of NBS for DRR
and CCA, particularly if these are carefully planned and managed, and
interconnected in a network of solutions (Debele et al., 2019; Kabisch et al.,
2016; Sahani et al., 2019). A nature-based approach to urban and peri-urban
development and management has been growing in popularity over the last
decade, but still needs to be fully integrated into national, regional, and local
policies. In particular, there is interest within the (re)insurance industry in
understanding the protective role of NBS in buffering risks posed by natural
hazards (Marchal et al., 2019).

Remaining knowledge gaps and a lack of comprehensive evidence on the


reversibility, flexibility, cost-effectiveness and feasibility, and/or long-term
sustainability of NBS as compared with grey approaches are barriers to
mainstreaming of NBS and their full incorporation within (re)insurance schemes
(Ruangpan et. al., 2020). This may deter decision-makers from investing in the
design and implementation of NBS for DRR and CCA rather than solely relying on
conventional grey solutions. Thus, additional NBS performance and impact data,
specifically evidence from field studies, is required to facilitate the integration of
these emerging concepts and NBS strategies in urban and regional planning and
design. The generation and dissemination of monitoring and evaluation data will
promote further NBS actions, creating a positive cycle for the generation of an
increasingly detailed knowledge base on NBS efficiency and cost-effectiveness
and informing the further development of policies regarding land management
and urban development (Kabisch et al., 2016).

1.2 NBS in European and International policy frameworks

1.2.1 NBS in the European policy context

To adapt to and mitigate the negative impacts of climate change and urbanisation
and to effectively address these challenges, decision-makers at local, regional
and global levels have gradually shifted paradigms away from a hard engineering
to a more adaptive and softer approach that enlarges the portfolio of options to
include NBS, including eco-engineering and ecological restoration. Since 2015,
within this new paradigm, NBS have been advocated by both policymakers and

20
practitioners as resilient, adaptable, resource efficient, locally adjustable, mainly
equitable, and optimised options to maximize opportunities to improve the well-
being of all urban residents, independent of their socioeconomic status, gender,
cultural background, or age (Faivre et al., 2017).

Nature-based solutions present a credible means to address key societal issues,


such as climate change, disaster risk, and biodiversity loss (SEP, 2021). A
multitude of scientific studies have demonstrated that NBS can contribute to
substantial improvements in air quality, microclimate conditions, and the health
and well-being of citizens. As such, NBS are highlighted in the European Green
Deal and recent key European policy initiatives, such as the EU Biodiversity
Strategy for 2030 (EC, 2020) and the new EU Strategy on Adaptation to Climate
Change (EC, 2021). In particular, the EU Biodiversity Strategy for 2030 highlights
the value and importance of NBS in fighting biodiversity loss, climate change and
other critical challenges, and promises funding for investment in NBS. Nature-
based solutions are also likely to play a key role in the new EU Forest Strategy
(currently under public consultation), and the forthcoming EU Soil Strategy and
European Zero Pollution Action Plan for air, water and soil.

The role of NBS as natural, functional infrastructure that can contribute to


sustainability, improve environmental quality and citizens’ well-being, whilst
simultaneously providing opportunities for economic development is consistent
with the EU Adaptation Strategy to climate change published in 2013 4 that aimed
to address climate adaptation in the European Union (EC, 2013). The strategy
specifically focused on enhancing the preparedness and capacity to respond to the
impacts of climate change at local, regional, national and EU levels, developing a
coherent approach and improving coordination (EC, 2013). The updated EU
Strategy on Adaptation to Climate Change issued in February 2021 5 specifically
highlights NBS as a cross-cutting priority area to support the further development
and implementation of climate adaptation strategies at all levels of governance
(EC, 2021). The EC has also expressed support for the ‘NbS for Climate Manifesto’,
proposed in August 2019 at the UN Climate Action Summit 2019.

NBS implementation can enhance the implementation of other major European


policies and strategies. Targeted NBS interventions are capable of enabling a
more comprehensive implementation of the Floods Directive 6 via complementing
national flood management strategies and flood risk management plans, e.g.
through natural flood management schemes; Groundwater Directive 7 via
interventions that reduce the burden on groundwater resources; and the Urban
Waste-Water Treatment Directive 8 via infiltrating a portion of surface runoff. The
overarching Water Framework Directive 9 enforces the implementation of the local
river basin management plans to which NBS contribute directly and indirectly.
Nature-based solutions contribute directly to the Water Framework Directive
(WFD) through integrated water management in terms of quality and quantity,

4
COM(2013) 216 final
5
COM/2021/82 final
6
OJ L 288, 6.11.2007, p. 27–34
7
OJ L 372, 27.12.2006, p. 19–31
8
OJ L 135, 30.5.1991, p. 40–52
9
OJ L 327, 22.12.2000, p. 1–73

21
which supports compliance with requirements for good ecological,
physicochemical, and other statuses of surface waters and groundwater set by
the WFD, as well as the active participation of stakeholders through co-design of
NBS measures for water security.

NBS for DRR strategies additionally contribute to the Marine Strategy Framework
Directive 10 via environmental targets and monitoring of coastal zones, the new
emphasis on the Blue economy, and indirectly to the EU Civil Protection
Mechanism by joint planning and coordination of disaster response activities for
enhanced prevention and preparedness to disasters. NBS employed for DRR
equally contribute to the Floods Directive by lessening the potential consequences
and magnitude of flooding at flood risk zones previously identified during the
preliminary flood risk assessment. The EU Action Plan on the Sendai Framework
for Disaster Risk Reduction (2015) builds on the Sendai Framework and the
associated international agreements and processes, to further enhance and
promote disaster risk management and its integration in EU policies. The EU
Action Plan on the Sendai Framework for Disaster Risk Reduction presents ways
that risks can be reduced through working with nature, while also providing
human, biodiversity and climate benefits 11.

Biodiversity emphasis, as the core of the NBS concept (cf. Section 1.1), observes
distinct ties with Natura2000 network, and the Birds 12- and Habitats 13 Directives
by directly re-establishing natural habitats and their connectivity, in compliance
with the EU goals on green infrastructure, reducing pressures on the local
biodiversity. The value of NBS for biodiversity enhancement in an urban
environment is outlined in the EU Green Infrastructure strategy 14.

NBS address the Air Quality Directive 15 via alleviating urban air pollution,
contributing to decreased local levels of particulate matter (PM2.5, PM10), nitrogen
dioxide (NO2) and ground-level ozone (O3) for protection of human health.
Explicitly addressing urban air pollution additionally contributes to the Clean Air
Programme for Europe 16.

Adaptation to the effects of climate change is equally reflected in the EU


Bioeconomy Strategy and the EU Circular Economy Strategy 17, both major
constituents of the European Green Deal 18. NBS can contribute to circularity by,
e.g., facilitating the recycling or productive re-use of organic materials, or
rainwater capture and re-use. The latter can significantly advance, for example,

10
OJ L 164, 25.6.2008, p. 19–40
11
https://ec.europa.eu/echo/sites/echo-site/files/sendai_swd_2016_205_0.pdf
12
OJ L 20, 26.1.2010, p. 7–25
13
OJ L 206, 22.7.1992, p. 7–50
14
COM/2013/0249 final
15
OJ L 152, 11.6.2008, p. 1–44
16
COM(2013) 918 final
17
COM/2020/98 final
18
COM/2019/640 final

22
the Water Scarcity and Droughts Policy 19, while helping to advance the EU circular
economy action plan 20 and approach through the water cycle

An abundance of EU legal acts ensures coordination within and across the policies
and strategies, all aiming at strengthening regional development. Being
interlinked by their nature, the Water Framework Directive itself encompasses
the links to the EU climate change strategy and other policies, such as those
related to agriculture (e.g., EU Common Agricultural Policy 21) and green
infrastructure. Strosser et al. (2015) remark that stakeholder participation and
awareness raising, which NBS influence directly, contributes to a more successful
implementation of the strategies outlined in the Directives. NBS projects, being
participatory in their nature, directly influence the Open Science initiative
established by the EU (EC, 2016) enabling education, research, and data-
informed decision- and policymaking.

The EU Research and Innovation (R&I) policy agenda on NBS and Re-Naturing
Cities aims to position the EU as leader in ‘Innovating with nature’ for more
sustainable and resilient societies. The main goals of this EU policy agenda are
to: (1) Enhance the framework conditions for NBS at EU policy level; (2) Develop
an EU Research and Innovation Community for NBS; (3) Provide the evidence
and knowledge base for NBS; (4) Advance the development, uptake and upscale
of innovative NBS; and (5) Mainstream NBS within the international agenda. This
agenda contributes to knowledge creation and policy development in relevant
areas, such as biodiversity, water management, climate change mitigation and
adaptation, sustainable development, and disaster risk reduction (EC, 2014; EC,
2020). This agenda proposes NBS as more effective and efficient solutions than
more traditional approaches – turning environmental, social and economic
challenges into innovation opportunities. At its core are the concepts of adressing
societal challenges with nature, accounting for and maximising multiple benefits,
co-creating and community building, establising an evidence base and
mainstreaming NBS in European and international policies. This handbook is the
result of work carried out by Horizon2020 NBS projects funded under the EU R&I
policy agenda.

1.2.2 NBS in an International policy context

Internationally, the Hyogo Framework for Action (HFA) 2005–2015 (UNISDR,


2005), is an international agreement under the auspices of the United Nations
International Strategy for Disaster Reduction (UNISDR), aimed to reduce the loss
of lives and damage to properties and overall economic impact from natural
hazards to enhance the sustainability of nations and communities (Quevauviller
and Gemmer, 2015). A lack of sufficient quantitative data necessary to evaluate
various options and actions to mitigate the impacts of natural hazards was
identified in the HFA. This lack of data has made monitoring the progress of
disaster risk reduction (DRR) and climate change adaptation (CCA) particularly
challenging (UNISDR, 2011). This Handbook contributes directly to the

19
COM(2007) 414 final
20
https://ec.europa.eu/environment/strategy/circular-economy-action-plan_en
21
OJ L 347, 20.12.2013, p. 549–607

23
acquisition of consistent and accurate data concerning impacts of actions
undertaken to address natural hazards in a systematic way.

In the new international policy agendas for DRR and CCA, founded on the Sendai
Framework for Disaster Risk Reduction (SFDRR) 2015-2030 (UNISDR, 2015) and
the Paris Agreement 22 on climate change, further effort was placed on more
effectively measuring DRR and CCA progress. In line with this, the 17 United
Nations (UN) Sustainable Development Goals (SDGs) identify a series of
objectives, clear targets and set of indicators to enhance, monitor, and evaluate
progress on environmental and human conditions (UN General Assembly, 2015).
Local monitoring of progress towards SDG achievement is strongly supported by
the impact evaluation framework presented herein. A number of the indicators
associated with SDGs have been adopted as part of the present framework and
are presented in this handbook and associated Appendix of Methods.

The HFA made little reference to nature or ecosystem-based approaches for DRR
and CCA compared with its successor, the SFDRR. This new frame was endorsed
by the UN General Assembly following the 2015 third UN World Conference on
DRR (WCDRR), re-enforcing the change in prevailing paradigm, with the clear
goal to build the resilience of nations and communities to disasters by shifting
towards disaster risk management and prevention. With the SFDRR agreement,
policy and decision-makers have committed to decrease global disaster damages
by 2030 and have recognized the key role of measuring disaster losses in
achieving this objective (UNISDR, 2015). The SFDRR has a global agenda in
reducing and averting disaster risks by reinforcing adaptation in society and
economic settings. It argues that DRR responsibility should be shared among the
different stakeholders including local government, the private sector, and others.
The SFDRR works in parallel with the other 2030 Agenda agreements, including
the Paris Climate Change Agreement, the Addis Ababa Action Agenda on
Financing for Development, the New Urban Agenda, and ultimately the 2030
Action Agenda for the SDGs. Many of these ambitious goals directly refer to the
urban and peri-urban environments where most of the global population live and
will increasingly expand in the future. However, as outlined, the impact of climate
change is extended to wider territories and actions since often adaptation
requires coordinated measures at a larger territorial scale.

Nature-based solutions can form a core element of local, regional, and national
policy initiatives. The need for a more “natural” living environment is increasingly
evident, with the importance of connecting with nature particularly recognised
during the COVID19 pandemic, primarily in urban and peri-urban areas, and the
public demands for greater attention to biodiversity and climate threats continue
to grow at the local and global scales. According to Langer (1995), in order to
achieve an ecological transformation of our economy and society, the process has
to be “socially desirable” for the majority of people. Thus, because of EU and
national level government incentives and directives and citizens’ requests, we are
living during a period of significant transition. Local governments and institutions
can employ this handbook as a tool to support the design and evaluation of NBS

22
United Nations Framework Convention on Climate Change. https://unfccc.int/

24
projects as part of the transition to a green, climate resilient and sustainable
society.

1.3 Purpose of the NBS Impact Evaluation Handbook

The need for robust methods, frameworks and indicators that allow the
quantification and the multiple levels of interaction associated to NBS, from co-
design to implementation is clear. This handbook provides a protocol for selection
of key indicators of NBS impact and methods for their assessment, which can be
applied to monitor reference parameters. The handbook adopts the EKLIPSE
Working Group impact evaluation framework approach with key challenge-based
indicators (Raymond et al., 2017). Building on the EKLIPSE framework, which
was primarily designed for urban areas, this handbook extends the original
EKLIPSE challenge areas to address additional challenges and scales of NBS
application (see Chapter 4 for details).

1.3.1 Handbook aim

This handbook offers an overall evaluation framework for NBS. It covers the
technical scope related to the monitoring processes relevant to stakeholders who
are involved in NBS assessment and implementation, such as the research
community, technology providers, authorities and NBS implementers. The
sequence of NBS evaluation framework development and implementation are
addressed from the conception-design and implementation of a monitoring and
evaluation plan through NBS monitoring and final evaluation of benefits and dis-
benefits. The indicators of NBS impact detailed within this handbook and the
accompanying Appendix of Methods encompass environmental, social, and
economic domains in the NBS assessment.

This handbook and its Appendix of Methods should be regarded as living


documents. Increases in scientific knowledge and the accumulation of evidence
on NBS performance and impact, together with technological advances, will
necessitate changes and updates to accommodate advances in the field of NBS
research. In addition, social and cultural change may alter how NBS are viewed
by decision-makers and the wider public, as well as the policy context within
which we view NBS. The authors anticipate periodic updates to this handbook to
account for changes to the scientific, technological, social, cultural, and political
landscape and the resultant impact on how we understand and use NBS.

In summary, the handbook serves as a comprehensive reference handbook,


based upon current best available knowledge and state-of-the-art technologies
and practices. It provides detailed information to guide the development and
implementation of an NBS monitoring and evaluation plan, and the use of the
NBS impact indicators presented as a query tool. This handbook contributes to
the provision of sustainable nature-based alternatives to environmental
challenges while addressing growing demands for the peaceful coexistence
between nature and humans (Sánchez et al., 2020).

25
1.3.2 Intended audience of this handbook

This handbook presents information in a way that aims to make NBS accessible
to educated non-experts, including all individuals and organisations interested in
NBS but primarily focused on the individuals and groups involved in creating,
implementing, and evaluating NBS. We focus on a “non-expert” audience because
NBS are capable of addressing numerous societal challenges while providing a
range of co-benefits across multiple expert domains. It is unlikely that a single
individual or even a single group of people will possess high-level expertise across
all domains addressed by NBS. Thus, this handbook aims to provide critical
background on the NBS concept and where it fits in a European and international
policy context, knowledge regarding the essential steps in developing and
implementing a monitoring and evaluation plan, guidance on the selection and
application of indicators of NBS impact, and knowledge of data to support
effective data management and use in NBS assessment.

The handbook, as an enabler of NBS knowledge, provides a user-friendly way to


plan, monitor and evaluate NBS. In this sense, the handbook functions a tool for
the main stakeholders of the NBS value chain to facilitate an improved
understanding of NBS impacts inform NBS implementation to address identified
concerns. In this sense, this handbook targets several NBS stakeholder groups,
including but not limited to:

• Policy makers, urban planners and other public agents involved in


urban development and land management. The handbook can aid the
development of coherent strategies for sustainable development, climate
change adaptation and mitigation, biodiversity enhancement, disaster
risk reduction, and a just transition and deep transformation towards
climate change resilience at both urban and regional scales. It can support
stakeholders in the development and implementation of NBS monitoring
and evaluation plans within the area of intervention as a tool assess the
achievement of specified objectives, thus providing valuable evidence of
NBS effectiveness and informing management actions.

• Members of the scientific community who wish to deepen their


knowledge on state-of-the-art tools and methods available for monitoring
progress towards specific, measurable environmental, social and
economic objectives, and to gather evidence regarding the provision of
ecosystem services (ESS) by NBS.

• Businesses, nature-based enterprises, impact investors, and


industries involved in the design, construction, and management of
NBS, or interested in the utilisation of or investing into the services that
NBS provide.

• Non-governmental organisations and civil society at large who are


interested in understanding the environmental, social, and economic
impacts of NBS and in gathering knowledge on the existing tools for
quantifying NBS impact will benefit from the comprehensive background
knowledge and detailed steps for key processes presented in this
handbook. In addition, this handbook provides information to support the

26
active engagement of citizens in the acquisition of data related to NBS
performance and impact through local monitoring programs, such as
citizen science or crowdsourcing of information.

1.3.3 How this handbook was developed

This handbook was developed by a large group of experts from several NBS-
related EU H2020 funded projects and European programmes to support the
development of a European evidence base on NBS performance and impact. Over
the past decade, the EC has adopted a series of strategies in response to the
challenges arising from anthropogenic pressures on the environment and
observed increases in natural hazards related to anthropogenic climate change.
Many of these strategies were focused on sustainable actions to mitigate the risks
derived from the human exposure to different kinds of threats. Specifically, from
2015 a large investment in research and development was made to improve
knowledge regarding NBS processes and functions, demonstrate their application
and derive evidence of NBS performance and impact across a range of different
application contexts. This translated into more than twenty H2020 projects and
programmes directly addressing the area of NBS and closely related themes,
including but not limited to (in alphabetical order): BiodivERsA, CLEARING
HOUSE, CLEVER Cities, CONNECTING Nature, EdiCitNet, EKLIPSE, GREEN
SURGE, GROW GREEN, Inspiration, MAES/EnRoute, NAIAD, Nature4Cities,
Naturvation, NetworkNature, OpenNESS, OPERAs, OPERANDUM, PHUSICOS,
proGIreg, RECONECT, REGREEN, Think Nature, TURaS, UNaLab, URBAN
GreenUP, and URBiNAT.

Table 1-1 illustrates the wide range of main objectives and expected outcomes
from these projects. The Projects range from those directly addressing the NBS
impact on climate change and water related issues in urban, rural and natural
areas, to others addressing the NBS impact on social cohesion, or links to the
insurance industry, and hydro-meteorological risks. More recently, project scopes
expanded to evaluating impacts on biodiversity and ecological restoration, and
collaborating with other global regions, such as China or Latin America. Several
web portals, networks, platforms and initiatives have been developed to address
NBS at European, national and sub-national levels. A non-exhaustive list of
networks, platforms and initiatives includes OPPLA 23, NetworkNature 24,
BiodivERsA 25, Biodiversity Information System for Europe (BISE) 26,
ThinkNature 27, the European Climate Adaptation Platform Climate-ADAPT 28,
Natural Water Retention Measures NWRM platform 29, and the EC Disaster Risk
Management Knowledge Centre (DRMKC) 30.

23
https://oppla.eu/
24
https://networknature.eu/
25
https://www.biodiversa.org/
26
https://biodiversity.europa.eu/
27
https://www.think-nature.eu/
28
https://climate-adapt.eea.europa.eu/
29
http://nwrm.eu/
30
https://drmkc.jrc.ec.europa.eu/

27
To integrate the outputs and promote the synergies emerging from these large
H2020 projects, several taskforces (TFs) were established linking the projects and
facilitating collaboration and knowledge exchange. These taskforces are comprised
of representatives from each of the H2020 NBS projects, representatives of the
Coordination and Support Action responsible for development and management of
the NBS Stakeholders platform, representatives from EASME and DG RTD, and
external observers from related programmes and initiatives. The six taskforces are:
TF1 - Data Management and EU NBS Knowledge Repository; TF2 - NBS Impact
Evaluation Framework; TF3 -: Governance, Business Models and Financial
Mechanisms; TF4 - NBS Communication; and TF6 - Co-creation for NBS. The
number of NBS taskforces and the focus of each will continue to evolve with time
as new needs are identified.

The present handbook was developed by members of TF2, whose collaborative


effort aimed at establishing a dynamic NBS impact evaluation framework based
on the collective experience acquired through execution of the NBS projects. One
of the primary goals of the taskforces is to jointly demonstrate the effectiveness
of NBS by providing a scientific evidence base detailing the performance and
impacts of NBS of different types as implemented in different contexts, and to
compile and disseminate best practices and guidelines for NBS development and
implementation based on participatory processes. Through concerted actions, like
this handbook, the taskforces are helping to define the framework to strengthen
NBS-based policies in accordance with local legislation, cultures and social norms,
while supporting new technologies and innovation in the area of NBS to promote
European leadership in the field.

28
Table 1-2. Summaries of previous and ongoing projects and programmes working on NBS (2007-2022).

Projects related to Aims, targets and brief summary Reference


NBS

BiodivERsA is a network of national and regional funding


organizations promoting pan-European research on biodiversity and
BiodivERsA http://www.biodiversa.org/
ecosystem services, and it is offering innovative opportunities for
the conservation and sustainable management of biodiversity.

CLEARING HOUSE is the first Sino-European research project on


urban forests and urban trees as nature-based solutions. We look
into how a traditional solution as urban trees can contribute to
CLEARING HOUSE http://clearinghouseproject.eu/
sustainable cities. The project aims to develop an online application,
a global benchmark tool, and guidelines to support the design,
governance and management of urban forests.

CLEVER Cities aims to increase and improve local knowledge of


nature-based solutions, demonstrate that greener cities work better
CLEVER Cities for people and communities, contribute data and information to EU https://clevercities.eu/
policy-making, and ultimately promote and enable the uptake of
nature-based solutions in urban planning world-wide.

CONNECTING Nature brings in actions to feed the initiation and


expansion of economic and social enterprises in production and
CONNECTING
large-scale implementation of NBS in urban settings to measure the https://connectingnature.eu/
Nature
impact of these initiatives on climate change adaptation, health and
well-being, social cohesion and sustainable economic development.

The Edible Cities Network focuses on Edible City Solutions, defined


as NBS related to urban food production, distribution and use.
EdiCitNet implements, monitors and transfers Edible City Solutions
EdiCitNet https://www.edicitnet.com/
in close cooperation with city authorities and other local
stakeholders. Thereby, it aims at increasing social, environmental
and economic sustainability of cities.

29
EKLIPSE aims to develop support mechanisms that facilitate
linkages between science, policy and society, through different
EKLPSE actions such as knowledge synthesis, identifying research priorities, http://www.eklipse-mechanism.eu/
and building the Network of Networks that will support the other
actions

EnRoute is a project of the European Commission in the framework


of the EU Biodiversity Strategy and the Green Infrastructure
Strategy. EnRoute provides scientific knowledge of how urban
EnRoute https://oppla.eu/groups/enroute
ecosystems can support urban planning at different stages of policy
and for various spatial scales and how to help policy-making for
sustainable cities.

GREEN SURGE prepared strategies to design urban green


approaches: integrating green and grey approaches, connecting
GREEN SURGE https://cordis.europa.eu/project/id/603567
green areas, utilizing the multipurpose character of the green
approach and involving citizens in urban planning.

GROW GREEN aims to invest in NBS (high-quality green spaces and


waterways) while long term city planning to develop climate and
GROW GREEN water resilience, strong and habitable cities, capable of dealing http://growgreenproject.eu/
major urban challenges, such as flooding, heat stress, drought, poor
air quality, unemployment and biodiversity-loss.

Imspiration aimed to develop a Strategic Research Agenda (SRA) to


inform environmentally friendly, socially acceptable and
economically affordable soil and land use management that meets
Inspiration societal needs and challenges. A SRA built on end-user knowledge http://www.inspiration-h2020.eu/
needs is more likely to be enthusiastically adopted by funders in
order to promote the knowledge creation, transfer and
implementation agenda.

The Working Group on Mapping and Assessment on Ecosystems and


their Services (MAES) was established under the Common https://ec.europa.eu/environment/nature/
MAES
Implementation Framework (CIF) to support the effective delivery knowledge/ecosystem_assessment/index_en.htm
of the EU Biodiversity Strategy to 2020. The objective of the MAES

30
Working Group is to provide guidance for the implementation of
Action 5 by the EU and its Member States, including development
of a coherent analytical framework to be applied by the EU and its
Member States in order to ensure consistent approaches are used
to map ecosystems and their services.

NAIAD is focused on developing a strong conceptual framework for


evaluating the assurance and the insurance value of ecosystem
services. The project has developed the concept of natural
assurance schemes, and the range of tools and methods to design
them, ranging from physical, social and economic assessment,
NAIAD http://naiad2020.eu/
integration and co-design with stakeholders, to the development of
business models and financing arrangements, and finally
implementation and monitoring. Stakeholders involved included
insurers, river basin agencies and local authorities, in the validation
and application in nine case study sites across Europe.

Nature4Cities aims for a positive balance between economic,


environmental and societal benefits and costs by creating a
reference platform for NBS, offering technical solutions, methods
Nature4Cities https://www.nature4cities.eu/
and tools for urban planning. This balance entails collaborative
models from citizens, researchers, policymakers and industry
leaders through co-creation processes.

NATURVATION assesses NBS achievements in cities, examines their


innovation process and works with communities and stakeholders
NATURVATION https://naturvation.eu/
to develop the knowledge and tools required for the recognition of
NBS potential for meeting urban sustainability goals.

NetworkNature is a European and global platform providing


resources for the nature-based solutions community and creating
opportunities for local, regional and international cooperation to
NetworkNature https://networknature.eu/
maximise the impact and mainstreaming of NBS. All interested
stakeholders can access and contribute cutting-edge, innovative
knowledge and expertise on NBS to the NetworkNature platform.

31
OpenNESS aims to translate the concepts of Natural Capital (NC)
and Ecosystem Services (ESS) into operational frameworks that
provide tested, practical and tailored solutions for integrating ESS
OpenNESS http://www.openness-project.eu/
into land, water and urban management and decision-making. It
examines how the concepts link to, and support, wider EU economic,
social and environmental policy initiatives.

OPERAs combined NBS with traditional engineered solutions by


constructing and maintaining semi-fixed dunes on Barcelona's
OPERAs https://www.operas-project.eu/
(Spain) urban coastline, aiming to optimize ecosystem benefits and
augment coastal defence against sea-level rise.

OPERANSUM is developing a set of co-designed, co-developed,


deployed, tested and demonstrated innovative NBS for the
management of the impact of hydro-meteorologial risks (HMRs),
OPERANDUM https://www.operandum-project.eu
especially focused in European rural and natural territories,
facilitating the adoption of new policies for the reduction of HMRs
via NBS and their promotion.

PHUSICOS is demonstrating the effectiveness of NBS and their


PHUSICOS ability to reduce the impacts from small, frequent events (extensive https://phusicos.eu
risks) in rural mountain landscapes.

proGIreg focuses on the implementation and observation of eight


different NBS for creating productive GI to improve living conditions
proGIreg and reduce vulnerability to climate change, and to provide www.progireg.eu
measurable economic benefits to citizens and entrepreneurs in post-
industrial urban districts.

RECONECT aims to rapidly enhance the European reference


framework on NBS for hydro-meteorological risk reduction by
RECONECT http://www.reconect.eu/
demonstrating, referencing, upscaling and exploiting large-scale
NBS in rural and natural areas.

32
REGREEN aims to substantially advance evidence and tools by
systematically modelling and combining ecosystem services and
biodiversity as the basis for urban NBS in Europe and China. This
also involves policy experimental learning, strategies for
REGREEN https://www.regreen-project.eu/
depavement, education and citizen science in schools, valuation of
benefits and costs and the development of business models for
realising spatially relevant NBS that provide multiple ecosystem
services and wellbeing.

ThinkNature developed a platform that supports the widespread


ThinkNature https://www.think-nature.eu/
understanding and the promotion of NBS.

TURaS offers examples of approaches for enhancing urban


TURaS sustainability, e.g., green walls that can be adopted in any location https://cordis.europa.eu/project/id/282834
and at an affordable cost.

UNaLab aims to develop a European Reference Framework on


benefits, cost-effectiveness, economic viability and replicability of
NBS by promoting smart, inclusive, resilient and sustainable urban
UNaLab https://unalab.eu/
communities through co-creation (with and for local stakeholders)
of Urban Living Lab (ULL), demonstrations, experiments and
evaluation of NBS for climate and water challenges.

URBAN GreenUP aims to develop, apply and validate a methodology


for Renaturing Urban Plans to mitigate the effects of climate change,
URBAN GreenUP https://www.urbangreenup.eu/
improve air quality, water management and increase the
sustainability of cities through innovative NBS.

URBiNAT focuses on the regeneration and integration of deprived


social housing districts. Interventions focus on the public space to
co-create with citizens new urban, social and nature-based relations
URBiNAT www.urbinat.eu
within and between different neighbourhoods. URBiNAT aims to co-
plan a healthy corridor as an innovative and flexible NBS, integrating
micro NBS emerging from community-driven design processes.

33
1.4 Content of this handbook

A wealth of scientific evidence demonstrates that NBS are capable of addressing


challenges across multiple environmental, social, cultural and economic
dimensions. In this handbook, we consider the impacts on the following 12 key
societal challenge areas (Figure 1-2):

1. Climate Resilience
2. Water Management
3. Natural and Climate Hazards
4. Green Space Management
5. Biodiversity Enhancement
6. Air Quality
7. Place Regeneration
8. Knowledge and Social Capacity Building for Sustainable Urban Transformation
9. Participatory Planning and Governance
10. Social Justice and Social Cohesion
11. Health and Wellbeing
12. New Economic Opportunities and Green Jobs

Figure 1-2. What are nature-based solutions (NBS)? (Image © European Union, 2021)

34
A core principle of NBS lies in responding to one or more societal challenges that
have been identified as a priority by the local community (IUCN, 2020). This
handbook provides additional information about each of the aforementioned 12
key societal challenge areas and how they can be addressed by NBS. One of the
most important concerns expressed by a large proportion of the world’s
population is climate change and its effects, including floods, droughts, heat
islands, biodiversity loss, and other impacts. Nature-based solutions are
increasingly viewed as a viable approach to sustainably address the negative
impacts of climate change, both in terms of climate change adaptation and
mitigation. Many urban areas can be uncomfortable to inhabit due to air and
water pollution, traffic and industrial noise, violence, and the impacts of climate
change-related extreme weather events. Some urban residents also cite concerns
regarding a lack of social cohesion, lack of physical activity and the absence of
nature 31. The introduction of green spaces and other types of NBS have been
shown to enhance urban liveability, for example, by reducing heat stress and
enhancing human thermal comfort (Majidi et al. 2019).

Monitoring and evaluation is essential to determine whether implemented NBS


respond effectively to the challenges identified. Chapter 2 of this handbook
describes the main principles guiding NBS performance and impact evaluation,
including the general steps in the development of a credible monitoring and
evaluation plan that is tailored to a specific local context. The NBS project life
cycle assessment is a cyclical process rather than sequential and may require
reassessing or changing plans at any point in the process (Figure 1-3).

Of these processes, the monitoring and impact evaluation phase of an NBS


project is based on a holistic approach capable of illustrating the wider benefits
and trade-offs of NBS and their impacts in a systematic way. The monitoring and
evaluation phase of the project comprises both observation (monitoring) and
analysis (assessment by stakeholders or end-users) of the respective NBS
project’s impact. Thus, the goal of monitoring and evaluation is to analyse,
interpret and document the outcome of an NBS project for use by policymakers,
stakeholders, and decision-makers at various levels. Chapter 3 of this handbook
outlines a stepwise approach to the development and implementation of an NBS
monitoring and evaluation plan. The chapter describes how to engage in a
structured process to connect strategic objectives with NBS actions and expected
outcomes. Chapter 3 presents a series of examples of innovative monitoring and
evaluation support tools developed in EC H2020 projects.

Monitoring and impact evaluation of NBS is supported by indicators of NBS


performance and impact, including biophysical, socio-economic and sustainability
indicators, which are targeted to the evaluation of specific aspects of NBS
effectiveness. Chapter 4 of this handbook describes the 12 categories of societal
challenges that can be addressed by NBS, and conceptually maps the 12
challenge areas against the UN Sustainable Development Goals. A series of
indicators to evaluate the performance and impact of NBS are presented in
Chapter 4, organised by challenge area and further separated into Recommended

31
http://www.euro.who.int/en/health-topics

35
and Additional indicators to support the development of a holistic monitoring and
evaluation scheme. The accompanying Appendix of Methods provides brief
descriptions of the techniques used to assess each indicator listed in the
handbook, and guides the implementation of selected indicators to assess NBS
performance and impact.

Figure 1-3. A schematic diagram showing the full life cycle of NBS such as monitoring and evaluation, cost-
benefit analysis (adapted from Kumar et al., 2020)

The Recommended indicators presented in Chapter 4 are considered the most


important ones to monitor NBS impact; however, the Additional indicators of NBS
impact can provide highly valuable information, depending on local context and
particular data needs. Chapter 5 presents several case studies from different
NBS projects illustrating the selection and application of both Recommended and
Additional NBS impact indicators. This chapter provides examples of how different
groups of indicators were selected to address specific questions. Each case study
presented in Chapter 5 includes a brief description of the NBS, the reasons for
the selection of specific indicators for that particular NBS and a brief overview of
how the indicators are applied and/or monitored. The case studies also describe
the stakeholders involved in co-design and co-monitoring of NBS and discuss the
barriers and lessons learned during or after NBS implementation, monitoring and
evaluation.

36
Chapter 6 of the handbook specifically addresses the implementation of NBS to
mitigate the impact of hydro-meteorological events, detailing experiences to date
and providing examples of NBS application for hydro-meteorological risk
reduction. This chapter begins with an overview of hydro-meteorological risk and
illustrates how a hybrid combination of NBS and technical engineering solutions
(green-grey solutions) can be particularly effective in DRR and natural assurance
or (re)insurance schemes. The case studies in Chapter 6 provide examples of
different combinations of indicators and assessment models that can be used to
evaluate the technical, physical, economic, social, and environmental
performance and impact of NBS implemented for DRR.

Monitoring and evaluation of NBS is based upon data. How do we know which
type of data is most appropriate, and the potential sources of data? The Appendix
of Methods briefly outlines the data required to determine each NBS impact
indicator listed in Chapter 4, and the case studies presented in Chapters 5 and 6
illustrate how the indicators have been applied to different NBS. Chapter 7
reviews the main types of data, sources of data and techniques used to generate
data for NBS monitoring and impact evaluation. This chapter is an important
resource during NBS monitoring and evaluation planning, as the content of
Chapter 7 aids the development of a robust, actionable plan for the collection,
management and use of data in NBS impact assessment.

1.5 Conclusions

In the face of current global challenges, particularly the need to adapt to and
mitigate climate change, it is essential that spatial and urban planning and
management find ways to effectively integrate climate action, from both the
mitigation and adaptation perspectives. Nature-based solutions integrate
knowledge and practices from numerous related concepts such as EbA, Eco-DRR,
LID, GI, SuDs, and WSUD with extensive stakeholder engagement through co-
creation, co-implementation, and co-management actions throughout the NBS
lifecycle. The capacity of NBS to deliver a broad range of environmental,
economic and social co-benefits is widely recognised by practitioners and policy-
makers alike, and increasingly highly valued by citizens themselves. Nature-
based solutions are a core element of European CCA and biodiversity strategies
(EC, 2020; EC, 2021). Nature-based solutions can also contribute substantially
to the achievement of the UN SDGs, particularly targets under SDG 11
Sustainable cities and communities. Whilst not explicitly mentioned in the Sendai
Framework for DRR, NBS can play a key role in disaster risk management and
prevention through the adoption of Eco-DRR strategies.

Robust evaluation of NBS performance and impact is essential to fully understand


their benefits and trade-offs. Monitoring and evaluation facilitates an
understanding of how NBS performance and impacts evolve with time, and
provides insights into their respective potential for up-scaling and replication
according to stakeholder needs and the local context (environmental, social, and
economic conditions). Major challenges for up-scaling and replication of NBS arise
from a lack of detailed and standardised monitoring methods, reporting protocols
and guidance at the different stages of the NBS life cycle.

37
The International Union for the Conservation of Nature (IUCN) recently released
standards for the design and assessment of NBS to support mainstreaming of
nature conservation and consistency of NBS application (IUCN 2020). Whilst the
IUCN standard does not cite definitive thresholds, it provides a systematic
framework to facilitate and support consistency in NBS design and assessment
based on solutions-oriented outcomes. This handbook is intended to provide
standardised methods of NBS monitoring and evaluation, reporting protocols, and
guidance based upon best practices learned during NBS project work. The NBS
impact evaluation framework, indicators and methods described in this handbook
are strongly aligned with the eight criteria and sub-indicators that comprise the
standard framework for NBS design and assessment defined by the IUCN (2020).

Monitoring and evaluation of NBS is essential, not only to measure the “success”
of individual NBS projects, but to inform further actions and provide evidence to
support effective land use planning and management, and policy-making. This
handbook serves a guide to developing and implementing an appropriate,
scientifically robust NBS monitoring and evaluation plan to support NBS
management to achieve targeted objectives as well as NBS replication and up-
scaling efforts. The generation and dissemination of monitoring and evaluation
data will promote further NBS actions, creating a positive cycle for the generation
of an increasingly detailed knowledge base on NBS efficiency and cost-
effectiveness in comparison with traditional grey approaches (Kabisch et al.
2016).

1.6. References

Debele, S.E., Kumar, P., Sahani, J., Marti-Cardona, B., Mickovski, S.B., Leo, L.S., Porcù, F., Bertini, F.,
Montesi, D., Vojinovic, Z., Di Sabatino, S., ‘Nature-based solutions for hydro-meteorological hazards:
Revised concepts, classification schemes and databases’, Environmental Research, Vol. 179, 2019,
pp. 1-20.

Dorst, H., van der Jagt, A., Raven, R., and Runhaar, H., ‘Urban greening through nature-based solutions –
Key characteristics of an emerging concept’, Sustainable Cities and Society, Vol. 49, 2019, 101620.

Eisenberg, B., Polcher, V., Chiesa, C., Nature Based Solutions Technical Handbook, D5.1 NBS Technical
Handbook, 31/05/18, Deliverable 5.1 for the UNaLab H2020 project, GA no. 730052, 2018.

Escobedo, F.J., Giannico, V., Jim, C.Y., Sanesi, G., and Lafortezza, R., ‘Urban forests, ecosystem services,
green infrastructure and nature-based solutions: Nexus or evolving metaphors?’, Urban Forestry and
Urban Greening, Vol. 37, 2019, pp. 3–12.

European Commission, A sustainable bioeconomy for Europe: strengthening the connection between economy,
society and the environment. Updated Bioeconomy Strategy, Publications Office of the European
Union, Luxembourg, 2018.

European Commission, Forging a climate-resilient Europe - the new EU Strategy on Adaptation to Climate
Change, COM/2021/82 final, Brussels, 24.2.2021.

European Commission, Nature-based solutions & re-naturing cities. Final report of the Horizon 2020 expert
group on ‘Nature-Based solutions and Re-Naturing cities’, Publications Office of the European Union,
Luxembourg, 2015.

European Commission, Nature-Based Solutions: State of the Art in EU-funded Projects, Publications Office of
the European Union, Luxembourg, 2015, 2020.

European Commission, Open Innovation, Open Science, Open to the World – A vision for Europe, Publications
Office of the European Union, Luxembourg, 2016.

38
European Commission Towards a research and innovation policy agenda for Nature-based Solutions and
Renaturing Cities https://op.europa.eu/en/publication-detail/-/publication/fb117980-d5aa-46df-
8edc-af367cddc202 2014

Faivre, N., Fritz, M., Freitas, T., de Boissezon, B., and Vandewoestijne, S., ‘Nature-Based Solutions in the EU:
Innovating with nature to address social, economic and environmental challenges’, Environmental
Research, Vol. 159, 2017, pp. 509–518.

Frantzeskaki, N., Bergström, S., Gorissen, L., Egermann, M. and Ehnert, F., ’Nature-Based Solutions
Accelerating Urban Sustainability Transitions in Cities: Lessons from Dresden, Genk and Stockholm
Cities’, Nature-Based Solutions to Climate Change Adaptation in Urban Areas. Theory and Practice of
Urban Sustainability Transitions, Cham, Springer, 2017, pp. 65-88.

International Union for the Conservation of Nature, IUCN Global Standard for Nature-based Solutions. A user-
friendly framework for the verification, design and scaling up of NbS. First Edition. Gland,
International Union for the Conservation of Nature, 2020.

Kabisch, N., Frantzeskaki, N., Pauleit, S., Naumann, S., Davis, M., Artmann, M., Haase, D., Knapp, S., Korn,
H., Stadler, J., Zaunberger, K., and Bonn, A., ‘Nature-based solutions to climate change mitigation
and adaptation in urban areas: perspectives on indicators, knowledge gaps, barriers, and opportunities
for action’, Ecology and Society, Vol. 21, No 2, 2016, p. 39.

Langer, A., Il viaggiatore leggero. Scritti (1961 – 1995), Sellerio, Palermo, 1995.

Majidi, A. N., Vojinovic, Z., Alves, A., Weesakul, S., Sánchez, A., Boogaard, F., and Kluck, J., ‘Planning Nature-
Based Solutions for Urban Flood Reduction and Thermal Comfort Enhancement’, Sustainability, Vol.
11, No 22, 2019, Art. no 6361.

Marchal, R., Piton, G., Lopez-Gunn, E., Zorrilla-Miras, P., Van der Keur, P., Dartée, K.W., Pengal, P., Matthews,
J.H., Tacnet, J.-M-, Graveline, N., Altamirano, M.A., Joyce, J., Nanu, F., Groza, I., Peña, K., Cokan,
B., Burke, S. And Moncoulon, D., ‘The (Re)Insurance Industry’s Roles in the Integration of Nature-
based Solutions for Prevention in Disaster Risk Reduction—Insights from a European Survey’,
Sustainability, Vol. 11, No 22, 2019, Art. No 6212.

‘Natural language: the latest attempt to brand green practices is better than it sounds’, Nature, 541, 2017,
pp. 133–134.

Quevauviller, P. and Gemmer, M., ‘EU and international policies for hydrometeorological risks: operational
aspects and link to climate action’, Advances in Climate Change Research, Vol. 6, 2015, pp. 74-79.

Raymond, C.M., Pam, B., Breil, M., Nita, M.R., Kabisch, N., de Bel, M., Enzi, V., Frantzeskaki, N., Geneletti,
D., Cardinaletti, M., Lovinger, L., Basnou, C., Monteiro, A., Robrecht, H., Sgrigna, G., Munari, L.,
Calfapietra, C., An Impact Evaluation Framework to Support Planning and Evaluation of Nature-based
Solutions Projects, An EKLIPSE Expert Working Group report, Centre for Ecology and Hydrology,
Wallingford, 2017.

Rockström, J., Steffen, W., Noone, K., Persson, Å., Chapin, F.S., III, Lambin, E., Lenton, T.M., Scheffer, M.,
Folke, C., Schellnhuber, H.J., and Nykvist, B., ‘Planetary boundaries: exploring the safe operating
space for humanity’, Ecology and society, Vol. 14, No 2, 2009, p. 32.

Ruangpan, L., Vojinovic, Z., Di Sabatino, S., Leo, L. S., Capobianco, V., Oen, A.M.P., McClain, M.E., and Lopez-
Gunn, E., ‘Nature-based solutions for hydro-meteorological risk reduction: a state-of-the-art review
of the research area’, Natural Hazards and Earth System Sciences, Vol. 20, 2020, pp. 243–270.

Sahani, J., Kumar, P., Debele, S., Spyrou, C., Loupis, M., Aragão, L., Porcù, F., Shah, M.A.R., Di Sabatino,
S., ‘Hydro-meteorological risk assessment methods and management by nature-based solutions’,
Science of the Total Environment, Vol. 696, 2019, Art. No 133936.

Science for Environment Policy (SEP), ‘The solution is in nature. Future Brief 24. Brief produced for the
European Commission DG Environment’, UWE Bristol Science Communication Unit, , Bristol, 2021.

Strosser, P., Delacámara, G., Hanus, A., Williams, H., and Jaritt, N., A guide to support the selection, design
and implementation of Natural Water Retention Measures in Europe - Capturing the multiple benefits
of nature-based solutions, Publications Office of the European Union, Luxembourg, 2015.

UNISDR, ‘Hyogo Framework for Action 2005-2015: Building the Resilience of Nations and Communities to
Disasters’, Geneva, United Nations Office for Disaster Risk Reduction, 2005.

UNISDR, ‘Hyogo Framework for Action 2005-2015 mid-term review’, Geneva, United Nations Office for
Disaster Risk Reduction, 2011.

UNISDR, ‘Sendai framework for disaster risk reduction 2015–2030’, Proceedings of the Third United Nations
World Conference on DRR, 14-18 March 2015, Sendai, Japan, United Nations Office for Disaster Risk
Reduction, 2015.

39
40
Drawing on knowlegde from projects
funded by the European Union

Nature4Cities
Nature Based Solutions for re-naturing cities:
knowledge diffusion and decision support platform
through new collaborative models
Ankara (TR) Alcala de Henares (ES) Metropolitan Milan (IT) Szeged (HU)

Designed for Policy makers & public urban planners, urban professionals and civil society
thebNature4Cities platform aims to provide them support at all stages of a NBS project. The
structure of the platform developped by the Nature4Cities Horizon 2020 project follows a
support framework made up of three stages presented here.

Image: Collective planting in Alacala de Henares, Photo © Mathilde Elie, Plante&Cité


SCOPE Technical solutions, methods and tools to empower urban planning decision
making and address the contemporary environmental, social and economic
challenges that European Cities are facing.

Create a NBS project Main Challenges addressed

Applying from the creation of an NBS project,


the platform offers users the challenge of choo-
1. Climate Resilience
sing the right approach to meet their needs. It is
proposed to use the NBS explorer to learn about 2. Water Management
specific NBS and their benefits and to work with
the project observatory to learn from realized 3. Natural and Climate Hazards
success projects. The pre-selection tool provi-
des support to select a specific NBS type that 4. Green Space Management
meets their urban challenges and constraints.
Finally, others tools based on satellite imagery 5. Biodiversity
analysis offer the possibility to diagnostic city
trends and to identify the best place to imple- 6. Air Quality
ment a specific NBS project.
7. Place Regeneration

Assess a NBS project 8. Knowledge and Social Capacity Building

The assessment stage aims to increase the chan- 9. Participatory Planning and Governance
ces to meet the initial goals. It consists of pro-
posing tools and methods to assess the impacts 10. Social Justice and Social Cohesion
of a NBS project for urban resilience and for the
environment and socio-economic features. A 11. Health and Wellbeing
simplified assessment of urban performance is
proposed to assess how a NBS can benefit its 12. New Economic Opportunities & Green Jobs
surroundings (insitu), a socio-economic assess-
ment is proposed to estimate the socio-economic
benefits, co-benefits and costs of a NBS project
and an environmental assessment can be used to Lessons learned
assess the impact of the NBS throughout its life
cycle (exsitu). Sharing the concerns about urban challenges
with a wide range of stakeholders and involving
them into the planning and decision process is
Implement a NBS project essential to garantee a fruitful incorporation of
NBS. This early inclusion is fundamental to anti-
Once a project is ready to be launched, the Na- cipate an NBS project by choosing an appropriate
ture4Cities platform also offers tools and met- NBS and selecting the best place to face the chal-
hods to build governance, financial and busi- lenges of the city given the urban context. This is
ness models, to involve citizens and to build the main contribution of the Nature4Cities plat-
inclusive projects. form which aims to provide the knowledge and
tools necessary for the design phase upstream of
the implementation of a NBS.

Municipal Administrations

Citizens

Scientists / Academia
Learn more
Planning experts about the project
www.nature4cities.eu
Green businesses
Learn more
about the platform
www.nature4cities-platform.eu

The Nature4Cities project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 73046
Drawing on knowlegde from projects
funded by the European Union

NATURVATION
NATure-based URban innoVATION
Newcastle (GB) Barcelona (ES) Utrecht (NL) Malmö (SE) Gyor (HU) Leipzig (DE)

NATure-based URban innoVATION is a 4-year project, funded by the European Commission and
involving 14 institutions across Europe in the fields of urban development, geography, innova-
tion studies and economics. Led by Durham University, NATURVATION’s partnership includes city
governments, non-governmental organisations and business. We will seek to develop our un-
derstanding of what nature-based solutions can achieve in cities, examine how innovation can
be fostered in this domain, and contribute to realising the potential of nature-based solutions for
responding to urban sustainability challenges by working with communities and stakeholders.

Photo © R. de Bruijn_Photography / Shutterstock.com


SCOPE Environmental and socio-economic benefits and
impacts of NBS innovation at city-scale.

Approach to Impact Assessment Main Challenges addressed

The Urban Nature Navigator (UNN) tool has


been developed by the NATURVATION research
1. Climate Resilience
project and it aims at enabling decision-ma-
kers to evaluate the contribution that NBS 2. Water Management
can make towards achieving their urban sus-
tainability goals. The tool includes a process 3. Natural and Climate Hazards
through which decision-makers can identify
their urban sustainability priorities or challen- 4. Green Space Management
ges and assess the contributions of six types
of NBS towards meeting those urban sustai- 5. Biodiversity
nability challenges. By providing insight into
how NBS contribute to various sustainability 6. Air Quality
goals, decision-makers may better understand
the multiple benefits of NBS and the trade-offs 7. Place Regeneration
involved when selecting between different in-
terventions or understand how NBS may con- 8. Knowledge and Social Capacity Building
tribute to the sustainability goals that are
prioritised by different stakeholders. 9. Participatory Planning and Governance

10. Social Justice and Social Cohesion


Involved Stakeholders and roles 11. Health and Wellbeing

The UNN was co-designed between November 12. New Economic Opportunities & Green Jobs
2016 and March 2019, involving project re-
search partners from Lund University, the Net-
herlands Environmental Assessment Agency,
Durham University, Utrecht University, Central Lessons learned
European University (Hungary) and Leibniz-In-
stitut für Länderkunde. During the develop- The interdisciplinarity approach applied to the
ment of the UNN, stakeholders from the project development of the UNN resulted in the creati-
partner cities (Barcelona, Gyor, Leipzig, Malmo, on of an impact assessment tool that integrates
Newcastle and Utrecht) have been consulted in environmental and socio-economic impacts and
the form of stakeholder workshops, dialogues can communicate potential benefits of NBS to-
and interviews. The numbers of stakeholders wards addressing various urban sustainability
who participated in the meetings varied from 8 challenges. The UNN is based on the inclusion
to 45, and the stakeholder involved had different of various indicators that according to the tool
professional backgrounds (e.g. urban planners, development approach are considered credi-
representatives from the public authorities, and ble, salient, legitimate and feasible to capture
members of local NGOs and community groups). the multi-dimensional benefits of NBS based
The results of the stakeholder consultation pro- on solid scientific evidence. This bestows the
cesses helped re-formulate certain aspects of the tool a credible character although the applica-
tool to be more user friendly. tion of such approaches should also reflect on
the importance of current practices outside the
academic fields. Although inter-and transdisci-
plinary processes can be challenging, the expe-
Municipal Administrations (FR/FL) rience from working with different disciplinary
researchers and stakeholders, highlighted the
Planning experts need for a clear planning process that include
iterative stages as well as for periods to integ-
Scientists / Academia rate stakeholder feedback.

Schools and kindergartens


Learn more
Consultant Companies and SMEs www.naturvation.eu

The Naturvation project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 730243
Drawing on knowlegde from projects
funded by the European Union

THINK NATURE
Platform for Nature-based Solutions
The main objective of ThinkNature project is the development of a multi-stakeholder communi-
cation platform that will support the understanding and the promotion of nature-based solutions
in local, regional, EU and International level. Through dialogue uptake facilitation and steering
mechanisms as well as knowledge capacity building, the ThinkNature Platform brings together
multi-disciplinary scientific expertise, policy, business, and society, as well as citizens. This plat-
form is fluent to use and attractive to a wide variety of actors and stakeholders because it merges
all aspects of NBS in a clear, pyramidal methodological approach. It creates a wide interactive
society that builds new knowledge with a wide geographical scope.

Image: Bucharest
SCOPE The objective of the ThinkNature project is the development of a platform that
supports the understanding and the promotion of Nature-Based Solutions (NBS).

Approach to Impact Assessment Main Challenges addressed

The ThinkNature platform is an integrated multi-sta-


keholder web solution designed to stir dialogue and 1. Climate Resilience
interaction on NBS through discussion forums and
debates in order to identify regulatory, economic and 2. Water Management
technical barriers and to communicate and promo-
te successful NBS (https://platform.think-nature. 3. Natural and Climate Hazards
eu/). The platform has private and publicly accessi-
ble sections. The public section of the platform is the 4. Green Space Management
NBS Knowledge Hub which includes NBS projects,
Case studies, Resources. The private section of the 5. Biodiversity
platform is used to foster the dialogue on issues re-
lated to NBS, stakeholder networking and has the 6. Air Quality
capability of hosting sections with restricted access.
Having all functionalities to share documents, tasks 7. Place Regeneration
and events, the users are encouraged to participate
in online brainstorming forums and debates and en- 8. Knowledge and Social Capacity Building
hance their knowledge of NBS. The networking hub
brings together multi-disciplinary NBS expertise. 9. Participatory Planning and Governance

10. Social Justice and Social Cohesion


Involved Stakeholders and roles
11. Health and Wellbeing
The ThinkNature strategy for stakeholder engage-
ment has a three-prong approach: 12. New Economic Opportunities & Green Jobs
1. Establishment of the regional stakeholder net-
works – Four regional ‘think and do’ tanks co-
vering the Mediterranean, Oceanic, Temperate
Continental and Northern Temperate regions Lessons learned
have been established with their respective
networks of local representatives. Wide communication of NBS is needed for both pu-
2. Brainstorming forums – ThinkNature organised blic administration units and citizens. Regarding ci-
two brainstorming forums to engage stakehol- tizens, enhancing public knowledge about NBS can
ders in the uptake of NBS at the regional and increase public awareness and affect the attitude of
local levels. More than 300 stakeholders par- citizens concerning these solutions, which can in-
ticipated. The forums made a significant con- fluence local decisions about NBS. In the context of
tribution to the science-policy-business-society enabling effective communication, technical infor-
multi-stakeholder dialogue on NBS. mation should be translated for the target groups.
3. Barrier landscape and policy analysis - Think- All available information should be localised and in-
Nature aimed to develop strategies to overco- terpreted so that impacts, and risks are easily un-
me existing barriers and decision-making hie- derstandable. As to impacts and focusing on mul-
rarchy coupled with the engagement of local tiple benefits, NBS provide a series of benefits and
stakeholders in addressing NBS as part of EU, support the handling of many global challenges.
regional and local strategies. This information should be disseminated to at least
all potential end users, which may be lead actors to
Municipal Administrations adopt new NBS. However, training regarding emer-
ging techniques is needed for planners, developers,
Citizen and construction professionals to make things hap-
pen. Towards wide-spreading NBS knowledge, net-
Scientists / Academia working can be crucial too. Specifically, the partici-
pation in networks, associations, and consortiums,
Green businesses which are linked to NBS approach, may contribute
to useful NBS knowledge acquisition.
Architects and Spatial Planners
Learn more
Construction and infrastructure companies www.think-nature.eu

The ThinkNature project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 730338
2 PRINCIPLES GUIDING NBS PERFORMANCE
AND IMPACT EVALUATION
Coordinating Lead author
Skodra, J.

Lead authors
Connop, S., Tacnet, J.-M., Van Cauwenbergh, N.

Contributing authors
Almassy, D., Baldacchini, C., Basco Carrera, L., Caitana, B., Cardinali, M., Feliu, E., Garcia, I.,
Garcia-Blanco, G., Jones, L., Kraus, F., Mahmoud, I., Maia, S., Morello, E., Pérez Lapeña, B.,
Pinter, L., Porcu, F., Reichborn-Kjennerud, K., Ruangpan, L., Rutzinger, M., Vojinovic, Z.

Summary

What is this chapter about?

In this chapter, you will learn the main principles guiding NBS performance and
impact evaluation. Good evaluation can be the basis for effective NBS
implementation, enable evidence-based policymaking, support policy learning
and facilitate flexible decision-making, via adaptive management, to ensure the
sustainable performance of NBS over time. Credible and appropriate impact
evaluation is based on scientific evidence and end-user experiences, is properly
scaled and is linked to policy directives.

47
First, we explain key terms such as performance, impact, monitoring and
evaluation (Section 2.1). Then, in Section 2.2, we describe the critical role of
performance and impact evaluation in supporting decision-making. In section 2.3
we respond the question: “How do you develop a credible and appropriate impact
evaluation?” We propose a set of general steps and principles necessary to
develop an NBS impact monitoring and evaluation (M&E) plan, and explain how
to tailor this plan to the specific type and size of an NBS in your local context.
Finally, we synthesise the issues related to the design of M&E plans based on
practitioners’ feedback from existing H2020 projects and provide several
examples.

How can I use this chapter in my work with NBS?

This chapter provides an overview of the general steps and principles that are
necessary to develop a credible impact monitoring and evaluation plan. The
challenges and knowledge gaps that may arise during the definition of a
monitoring and evaluation strategy are also explored in this chapter.

When should I use this knowledge in my work with NBS?

Chapter 2 should be used at the beginning of the planning process for NBS
monitoring and impact assessment. Timely planning enables allocation of the
necessary time and resources to develop and implement the impact evaluation
plan, identify potential data gaps, and address funding constraints. These
principles can be revisited after initiating NBS monitoring to ensure that all
relevant and applicable steps of the process are being deployed.

How does this chapter link with the other parts of the handbook?

Chapter 2 introduces practical steps and principles for impact evaluation of NBS
measures in urban and rural settings. The individual impact monitoring steps are
further elaborated in Chapter 3.

2.1 Introduction and definitions

Impact evaluation is part of a broader agenda of evidence-based policy-making


and is essential to building knowledge about the effectiveness of interventions by
highlighting what does and does not work to achieve desired change (Morton
2009). To achieve this, impact evaluation systematically and empirically
examines the causal effects of the change in the built or natural environment
associated with the NBS intervention. These effects can be grouped into 12
societal challenges 32 and often impact simultaneously across multiple dimensions
(e.g., Place regeneration and Health and Wellbeing). Thus, impact evaluation is
related to the interpretation of indicators selected to assess NBS performance

32
Climate resilience, water management, natural and climate hazards, green space management, biodiversity
enhancement, air quality, place regeneration, knowledge and social capacity building for sustainable
urban transformation, participatory planning and governance, social justice and social cohesion, health
and wellbeing, new economic opportunities and green jobs (see Chapter 4).

48
and effectiveness in addressing challenges and fulfilling objectives. The main aim
of the impact evaluation is to answer a particular cause-and-effect question:

What is the impact (or causal effect) of an NBS intervention on an outcome of


interest?
It is therefore essential to define in advance what impacts (or effects) an NBS
intervention is expected to have, so that appropriate data at the appropriate scale
(e.g., spatial and temporal) may be collected (Morton, 2009). Meaningful impact
evaluation appropriately represents the NBS intervention in question and its
context. It should be valid in all respects (e.g., providing for both internal and
external validity 33) and provide useful information that can help inform future
directions. In order to understand why aspects of an intervention worked or did
not work, additional information on characteristics of NBS intervention are
necessary to understand the reasons for effectiveness (Morton, 2009) and the
conditions necessary for replicating the results in different context. In that sense,
significant support from monitoring is essential to complement the impact
evaluation.

The main characteristics of monitoring and evaluation are described in the


following paragraphs to enable differentiation between different approaches
suitable for NBS impact assessment.

Monitoring is a continuous process that tracks:

• The implementation process in order to determine what takes place and


when, during a project. The collected data are used to inform project
implementation, day-to-day management (adaptive management,
management of risk) and decisions related to effective implementation
processes and governance, and addressing challenges associated with
these processes.

• NBS performance against expected results (related to 12 societal


challenges3) and compared with measurements of a reference situation
(baseline). NBS performance is defined as the degree to which NBS
address an identified challenge3 and/or fulfil a specified objective in a
specific place (territory), time and socio‐economic context (Raymond et
al., 2017). It measures:

1. Change towards certain targets* (in this case performance thresholds


must be set - targets bring an additional challenge relating to how they
are selected /set); or ,

2. The change in relation to the Baseline/Reference; or,

3. A combination of numbers 1 and 2.

33
Internal validity refers to study design (factors like selection bias, spillovers, etc. should be addressed) and
external validity refers to generalizability (applicability of lessons-learned to another context or conditions)

49
Performance can be assessed by comparing against results from before the
intervention, from different NBS interventions or from alternative non-NBS
interventions, and may also analyse trends over time. The collected (qualitative
and quantitative) data is used to assess Key Performance Indicators (KPIs)
needed in impact evaluations.

Monitoring is therefore a critical source of information about NBS performance


(e.g., in terms of effectiveness, see Figure 2-1), including implementation and
costs, which supports the evidence base for both new and existing NBS.
Monitoring is used to reflect the reference situation before/without NBS and the
situation after/with the NBS implementation. In order to generate the most
relevant data from this process, monitoring should be conducted at an
appropriate scale taking into consideration urban morphology and regional
characteristics. A range of stakeholders may be involved in the local monitoring
teams, in different forms of participation - from informative to co-monitoring
activities.

Establishing a common standard for key indicators is important for comparing


NBS effectiveness across cities or regions. This helps to make results transferable
and thus support decision-makers in demonstrably effective and evidence-based
design of interventions in the built environment as well as in the natural
environment.

Evaluation is periodic, objective (un-biased, well-documented) assessment of a


planned, ongoing, or completed NBS project used selectively to answer specific
questions related to design, implementation, and results. It should be conducted
at the appropriate scale (e.g., spatial and temporal) according to different
decision-making contexts. In general, evaluations can address three types of
questions (Morra Imas and Rist, 2009):

• Descriptive questions explore what is taking place related to conditions,


processes and stakeholder views;

• Normative rating questions assess ‘what is’ taking place in comparison to


‘what should be’ taking place and apply to inputs, activities and outputs;

• Cause-and-effect questions explore what difference the NBS intervention


makes to outcomes.

Impact evaluation mostly addresses the cause-and-effect questions. The basic


evaluation question - what is the causal effect (impact) of an NBS intervention
on an outcome of interest? – can be applied to different contexts. For example,
what is the impact of the NBS on the mitigation of the adverse effects of hydro-
meteorological risks (that at the same time deliver socio-economic and well-being
benefits)? What is the impact of the residents’ participation in the NBS co-creation
on the use of the NBS, social cohesion and human health and well-being aspects?
How can broadening the scope of the evaluation of NBS projects engage diverse
funding sources necessary for city-wide implementation of NBS?

In that sense, impact evaluation focuses on the attribution and causality. To be


able to establish the causal effect and to attribute it to the NBS intervention

50
different methods can be used. These methods should estimate what the outcome
would have been for the area and for its users (residents, people working in that
area, etc.) if the NBS had not been developed (Morton, 2009). Alternatively, is a
given NBS intervention effective compared to the absence of the intervention or
to alternative, traditional engineering or planning solution? According to the
causality view, X (NBS intervention) causes Y (an outcome, e.g., alters
microclimate or social cohesion) and without X, Y would not exist.

Why are measurements needed in reference areas with no intervention?

Impact evaluation should use appropriate methods to prove that an NBS


intervention (X), rather than other changes in environment, society, etc. - has
caused a specific outcome (Y). However, NBS full development and changes in
the built environment usually take a longer period of time, during which other
factors may change as well. Thus, a whole range of effects can occur in the
meantime, that may change the behaviour and perception of the population but
have nothing to do with the original NBS intervention. This can be a global crisis
(such as the Corona pandemic), but also local events (such as particularly mild
weather for a longer period of time or a good score in sports events) that may
change the feeling of happiness of the population independently of the original
intervention.

One of the methods to filter out these effects, to prove the causality (Morton,
2009) and be able to attribute the outcome to the NBS intervention is a
comparison 34 of the treated area (NBS implemented) with a control area that has
not received a treatment (no NBS implemented). If an outcome of interest, e.g.
microclimate or social cohesion, has improved in both areas it means that there
were other factors that caused that change, rather than the NBS intervention. In
cases where an outcome of interest, microclimate or social cohesion, has
improved only in the treated area, then that change can be attributed to the NBS
intervention.

Treated and control area are assessed before (pre) and after (post-) -the NBS
intervention. The main challenge is to identify a control area and construct
population group that is as similar as possible to the treated area/group and be
in time before the participation and implementation process begins. In that sense,
timely planning of impact evaluation will enable allocation of the necessary time
and resources, and minimise funding constraints.

The definition of suitable “control area/group” or “before/after status” may not


be applicable in all cases, for example, where NBS are designed to mitigate
hydro-meteorological risks with relatively long (>10 years) return periods, such
as floods and droughts (see Chapter 6). Under such a scenario, modelling could
be an option, or evaluation of the impact of NBS on less severe (and more
frequent) events.

34
Example of a comparison to determine the impact of a programme or policy
https://ec.europa.eu/jrc/en/research-topic/counterfactual-impact-evaluation

51
For certain impact assessments of large-scale NBS, finding a suitable control area
can be challenging. Ideally, the control area should have similar environmental
and socio-economic conditions as the treated area but be located far enough to
be unaffected by the NBS intervention (to avoid spillover effect). If no suitable
control area can be identified, an alternative approach may be to predict what
the situation would be in the project area without implementation of the NBS.
This would become the reference situation to which post-NBS monitoring data
could be compared to assess the impact of NBS.

2.1.1 The concept of effectiveness

NBS effectiveness is defined as:

the degree to which objectives are achieved and the extent to which
targeted problems are solved. In contrast to efficiency, effectiveness is
determined without reference to costs (Raymond et al., 2017, p. vi).

For example (based on Raymond et al., 2017):

• Does the NBS lead to enhanced climate resilience in the urban area?

• Does the NBS lead to environmental benefits?

• Does the NBS lead to social benefits?

• Does the NBS lead to economic benefits?

• Does the NBS lead to biodiversity benefits?

In cases when NBS interventions combine solutions to achieve different impacts,


it is important to ensure that the impacts and its cumulative effects are integrated
throughout the process rather than simply synthesised at the end (Morton 2009).
This makes the whole analysis of their effects and impacts complex, increasing
uncertainty with respect to data collection.

A functional analysis using safety and reliability analysis concepts (Figure 2-1)
can help identifying the different system’s components, their functions, their
objectives and therefore their effectiveness. This methodology, classically used
for technological systems is innovative and helpful to model the whole system
and the interactions, as well as to break down the protected system into
components with given functions. The concept of components’ function and
corresponding objectives identification is key to design and choose the best
indicators for each application context. For example, a soakaway designed to
divert road drainage can also be planted with shrubs and other plants to support
pollinators. In that case, it is necessary to not only select indicators that measure
the quantity of drainage waters diverted or extent of flooding avoided, but also
indicators related to numbers of pollinators visiting flowers, etc. However, it is
essential to avoid overlapping indicators in the projects' framework. Clustering of
indicators can be handy for NBS effectiveness comparisons across cities or
regions and help decision-makers to move towards better solutions.

52
Based on the project objectives the assessment of the performance and the
effectiveness of a particular NBS intervention should take into account spatial and
temporal scale as well as specific target groups. Important part of impact
evaluations is an assessment of cost-benefit or cost-effectiveness. Knowing
which NBS interventions are effective and at what cost is crucial for
informing decisions about whether an intervention could be scaled up
and replicated.

Figure 2-1. Effectiveness indicators are designed to measure the extent to which NBS capacity reaches the
objective linked to an explicitly identified function (adapted from Tacnet et al., 2021)

Since benefits do not only refer to the physical sphere but include
social/individual, economic, and ecological/environmental benefits as well, the
complementary use of several evaluation approaches such as ex ante
simulations, mixed method analysis (drawing on both qualitative and quantitative
data), modelling and process evaluations can complement impact evaluations. It
is therefore important to note that there are always alternative approaches to
assess benefits, including those, which are non-monetisable. For a customised
impact assessment, it may therefore be helpful to adapt methods to one another
(e.g., by adding other dimensions to an already planned questionnaire) in order
to arrive at an effective impact assessment. In addition, integrating assessment
methods such as multi-criteria analysis or natural capital evaluation methods can
be adopted.

53
2.2 Decision-making context and impact evaluations: from needs
to indicators

This section provides a broad vision of decision-making contexts explaining why


NBS impact evaluations are needed. The aim is to identify and describe the
evaluation needs in general, independent of a specific project or objective.

Impact evaluation focuses on results of NBS interventions and provides a set of


tools that stakeholders can use to verify and improve the quality, efficiency, and
effectiveness of the interventions at various stages of implementation. Although
impact evaluation is a core driver of decision-making, since it is resource (time
and expertise) demanding it can remain a marginal activity. In that sense, it is
important that impact evaluation is designed at the early planning phases of an
NBS intervention, in order to allocate necessary resources, develop the
stakeholder engagement strategy and, where possible, integrate citizen science
in the design of the evaluation. Additionally, it is important that its value is
thoroughly communicated in order to support appropriate mainstreaming and
management.

In general, there are two main approaches to NBS impact evaluation:

1. NBS has already been developed in the past and the main aim is to
determine whether the NBS intervention is effective (retrospective impact
evaluation, i.e., ex-post evaluation). If NBS is already there and baseline
data was not collected before the NBS was implemented, it is difficult to
analyse whether the NBS is successfully implemented and whether the
envisioned outcomes are achieved (challenges related to the selection of
appropriate treated and control groups before the implementation).
However, this can be done for specific indicators using data that was
collected during the monitoring of the NBS and data collected for other
purposes (e.g., regional statistics of city administration data).

2. NBS has to be chosen during the planning phase (in comparison to


alternative solutions or business-as-usual, i.e., ex-ante evaluation
including screening) and implemented. Impact evaluations are developed
at the same time as the NBS intervention is being planned and are
integrated into the NBS implementation (prospective impact evaluation,
i.e., ex-ante evaluation including screening). Baseline data are collected
before the NBS intervention is implemented for both the area and/or group
receiving the intervention (the treated area/group) and the area/group
used for comparison that is not receiving the intervention (the control
area/group).

In both cases, the robust evidence generated by impact evaluations is important


for greater accountability, innovation, and learning in a decision-making context.
Learning and innovation demand a willingness to take risks and experiment.
Interdisciplinary nature of impact evaluation can contribute to busting
departmental silos and understanding broader benefits and co-benefits of NBS.
The accountability is crucial when it comes to reporting to funders, influencing
decision-makers and engaging novel funding streams (Gertler et al., 2016).

54
In that sense impact evaluations should provide credible evidence on
performance of the NBS and on whether a particular NBS intervention has
achieved or is achieving its envisioned outcomes. Impact evaluations require the
interpretation of those indicators that have been chosen to assess the benefits
and co-benefits over a period of time. In this respect, an important challenge is
how to look at the different indicators as a whole, considering their variation at
different time scales. It is also necessary to decide in advance how large an effect
is desirable and establish thresholds of impact. This is required in order to design
an evaluation with the appropriate degree of statistical power to be able to detect
an effect of the size expected. However, it is important to avoid a situation
whereby even a smallest change is interpreted as a success or failure of the NBS
(Gertler et al., 2016).

The question concerning uncertainty and more generally information imperfection


is very important here. Information imperfection (including uncertainty) can
apply to data features (e.g., resolution, coverage/spatial extent, etc.) and come
from type and reliability of sources (number of monitoring locations, experts) and
also from the evaluation procedure, measurement method or model themselves.
This is an important aspect as it carries the weight and reliability of
recommendations that will come from the monitoring and evaluation work. In
that sense, it is recommended to assess and propagate information quality during
the process of evaluation. The risk of failure of the monitoring system requires
the development of protocols to adopt mitigation measures in case a failure in
the monitoring system is detected.

In the decision-making context, the ability to replicate results is fundamental to


questions about the broader effectiveness and scalability of a particular NBS. In
addition to assessing the effectiveness of NBS in terms of desirable outcomes, it
is important to carefully trace a theory of change 35 that explains the process
through which NBS intervention has achieved the final outcome (benefits, co-
benefits, but also unintended negative effects). As illustrated in Figure 2-2, the
process begins with determining the desired long-term impacts related to the
project objectives/challenges (vision). Proceeding from the identification of the
existing conditions (reality), the necessary inputs and outputs are identified to
achieve short-term as well as intermediate outcomes, which themselves lead to
the desired long-term impact (vision). Assumptions identify the locally specific
risks and conditions that are present in the project’s context and attempt to
manage these risks by identifying what conditions must hold true for change to
occur. Understanding the process through which the changes have been
implemented enables the identification of causal pathways (Morton, 2009),
explaining:

• how the development of NBS functions in producing outputs, and

• how the process of producing outputs influences the final outcome.

35
A theory of change is a description of how an intervention is intended to deliver the desired results. It
describes the causal logic of how and why a particular program or intervention will reach its intended
outcomes. A theory of change is a key underpinning of any impact evaluation, given the cause-and-effect
focus of the research (Gertler et al., 2016, p. 32).

55
Figure 2-2. Example of the Theory of Change
(simplified adapted from The Young Foundation, CLEVER Cities project - D4.3/ WP4, pp. 18)

In order to gain a full picture of results, it is necessary to combine impact


evaluations with monitoring and complementary evaluation approaches (i.e., to
determine was the NBS implemented as planned, to provide context and
explanations to quantitative analysis – qualitative data and mixed methods 36).
Moreover, in the decision-making context a long-term, transdisciplinary studies
that focus on comparisons between NBS and non-NBS alternatives are very
valuable to policy-makers (Dick et al., 2020).

NBS are always implemented to fulfil a range of specified functions (e.g., reducing
floods, reducing air temperature, etc.), which can relate either to a quantifiable
parameters (e.g., water storage volume) or to a qualitative metric such as an
index to assess the well-being of a population.

In practice, assessing NBS’ effectiveness can be seen as several decision-making


problems:

a) Choosing - what is the most effective NBS?

b) Sorting - to which category of effectiveness or impact (low, medium, or


high) does the NBS belong?

c) Ranking - what is the effectiveness of NBS ranking from the worst to the
best (or vice versa)?

Multi-criteria decision analysis (MCDA) 37 is a way to gather any kind of qualitative


and quantitative criteria, which correspond to NBS impacts (Figure 2-3; see
Langemeyer et al., 2020; Harrison et al., 2017).

36
Mixed methods – an expert or a team of experts from different disciplines seeks to integrate quantitative
and qualitative approaches to theory, data collection, data analysis and interpretation. The purpose is to
strengthen the reliability of data, validity of the findings and recommendations, and to broaden and deepen
our understanding of the processes through which program outcomes and impacts are achieved, and how
these are affected by the local context. (Bamberger, 2012)
37
More information on multi-criteria decision analysis (MCDA), PP.129-139

56
Figure 2-3. The analysis of the effectiveness or impact of NBS can be done through a combination of
decision-aiding approaches and thematic, expert analysis and indicators. Features related to impact (effects)
of NBS are combined in a multicriteria decision-making framework including technical (T), organisational (O)
– not represented, physical (P), human (H), economic (E) and Environmental (E) considerations (TOPHEE
framework) (Tacnet et al., 2021, based on the NAIAD project D5.4).

In practice, those criteria can be linked to measurable indicators coming from


thematic, expert analysis. An interesting point is that it is a multidisciplinary
framework, which can easily link deterministic, physical assessments and a global
aggregated model as shown in Figure 2-3. In addition, this allows differentiation
between factual, objective assessment and more subjective evaluation based on
decision-makers’ preferences.

Planning frameworks move proactively towards adaptive planning and


management models, as a response to uncertainty and as an option to effectively
harness resilience (adapted from IUCN, 2020 38). In this context, it is imperative
that NBS implementation includes provisions to enable this adaptive planning and
management, generating evidence-base provided by regular monitoring and
evaluation, drawing on local knowledge as well as on scientific understanding.
NBS effectiveness and continuous performance evaluation are relevant
throughout the life-cycle of the intervention for identifying deviations, maximizing
synergies and total impacts, assessing and mitigating potential trade-offs, and
minimizing stranded investments.

38
https://www.iucn.org/theme/nature-based-solutions/resources/iucn-global-standard-nbs

57
2.3 Principles for the development of impact monitoring and
evaluation plans

Since evaluation plans are developed to evaluate benefits, co-benefits, and


negative effects as well as to evaluate performance of NBS in achieving
predefined objectives, this may require combining results of several impact
evaluations (each requiring its individual impact evaluation plan). The first section
lists general steps in designing and implementing an impact evaluation plan
(Figure 2-4). The second section presents main principles that should be followed
when developing steps of impact evaluations plans (Figure 2-4).

Figure 2-4. General steps and main principles involved in the development and implementation of an
impact evaluation plan.

2.3.1 Steps

The design of an impact evaluation plan is a multi-faceted process. Based on the


literature review and existing NBS projects we list six steps for developing impact
monitoring and evaluation plans. This is a general overview that will be explained
in more detail in Chapter 3.

STEP 1: Constructing and adopting a theory of change (Figure 2-2), which helps
to identify objectives and challenges, as well as outlining the process for achieving
the intended outcomes and impacts.

STEP 2: Developing a results chain to outline the theory of change – this covers
both the implementation process and the results outcomes.

STEP 3: Specifying the evaluation question(s), the basic impact evaluation


question is ‘What is the impact (or causal effect) of an NBS intervention on an

58
outcome of interest?’ The focus is on the Impact - the changes directly
attributable to an NBS intervention.

STEP 4: Selecting indicators and gathering data that answer the evaluation
question(s) and that allow the assessment of performance and process: ‘Does
NBS operate as designed and is it consistent with the planned theory of change?’
Critical selection of indicators that will be used to measure success/effectiveness
of the NBS intervention, as well as cause-and-effect indicators should focus the
evaluation, establish link to interventions well-defined objectives and assure that
outcome is attributable to the NBS.

STEP 5: Implementing the impact evaluation, evaluating positive/negative


features of NBS impacts related to the different challenges 39, analysing and
interpreting the findings.

STEP 6: Disseminating results and achieving policy impact

2.3.2 Principles

A proper assessment and evaluation of the targeted impacts is needed in a way


that is relevant and useful firstly to immediate end users and secondly to inform
broader policy processes. Therefore, development of impact monitoring and
evaluation plans should consider a few universal principles. Impact evaluation
plans and its indicators must:

1. Be scientifically sound,

2. Be practical and straight-forward,

3. Use reference conditions and baseline assessment,

4. Align with policy principles and reporting obligations,

5. Be based on a transdisciplinary approach.

These principles are explained below. Examples of the implementation of these


principles can be found in the selected NBS project example boxes between each
chapter.

39
In this Handbook impacts of nature-based solutions are assessed across 12 societal challenge areas: Climate
Resilience; Water Management; Natural and Climate Hazards; Green Space Management; Biodiversity;
Air Quality; Place Regeneration; Knowledge and Social Capacity Building for Sustainable Urban
Transformation; Participatory Planning and Governance; Social Justice and Social Cohesion; Health and
Well-being; New Economic Opportunities and Green Jobs – see Chapter 4

59
1) Impact evaluation should be scientifically sound

Since impact evaluations measure the change in an outcome that is attributable


to a defined NBS intervention, it is based on models of cause-and-effect. It
requires a credible and rigorously defined study design to control for factors other
than the intervention. However, cause-effects are not necessarily the only model.
In cases when the purpose of impact evaluation is raising awareness of the impact
of the NBS, the crucial factor is engagement of communities and decision-makers.
In that case, attribution may be replaced with contribution analysis 40. Ideally, in
a Theory of Change, aspects such as ‘community engagement’ can also be
assessed to demonstrate success of the project.

Measuring the impact of an NBS intervention should follow a concrete selection


of appropriate methodology that is capable of assessing the Key Performance
Indicators (or KPIs). Quantification and assessment of indicators is needed for
every challenge (environmental, economic, social or other4). But how to select or
develop indicators to be scientifically sound? This handbook provides an extended
list of scientifically sound indicators (Chapter 4) and examples of their application
(Chapter 5). The accompanying Appendix of Methods provides full descriptions of
each indicator and provides a brief methodology for each.

In case further indicators are necessary, based on a scientific literature the


following criteria can be used for their development (Figure 2-5):

Figure 2-5. Criteria for developing ecosystem service indicators


(adapted from Van Oudenhoven et al., 2018)

40
Contribution Analysis is a structured approach that enables assessing real-world challenges. It consists of a
step-wise, iterative process of refining Theory of Change. It does not seek to conclusively prove whether,
or how far, a development intervention has contributed to a change. Instead it seeks to reduce
uncertainty (https://www.intrac.org/wpcms/wp-content/uploads/2017/01/Contribution-analysis.pdf).

60
1. Credibility: the process of indicator development should be based on a
review of existing literature and on an external review by experts,
controlled path of production, elaboration, validation and monitoring of
data according to scientific protocols and methodologies: scientific
selection methods, validation, integration into methodology, triangulation
of data.

2. Salience: relates to the capacity of indicators to convey useful and relevant


information for decision makers about specific objectives as perceived by
potential end-users and stakeholders. It is important to use effective
means to present and translate scientific indicators in a way that it is easy
to communicate to non-experts: easy to read, understandable and not
generating misunderstanding (visualisation, modelling and simulation
tools: such as graphical, GIS, tabular, model animations, landscape design
drawings, etc.). Indicators should be temporary explicit to have the
potential to monitor change and assess progress over time. Moreover,
indicators should be scalable and transferable.

3. Legitimacy: selection on the basis of relevant indicators to meet the scopes


of monitoring process (for example, SMART 41): the selection of the most
appropriate model of impact evaluation will depend mainly on vision and
outcomes of interest in the project, scale of implementation, desired co-
benefits and available resources allocated to monitoring work and time.
The impact monitoring and evaluation plans need to be iterated and co-
produced with the relevant stakeholders and experts from different
disciplines (see principle 5 on transdisciplinarity) and not be a one-way
communication or design. In addition, indicators should be the outcome of
a shared process, to meet the expectations of a wide number of
stakeholders and, where possible, to express the engagement of
communities in decision-making and raise the awareness.

4. Feasibility: relates to the sufficiency of data, time and resources to assess


and monitor indicators (simple indicators are easy to acquire, easy to
elaborate, assess, and monitor over time). Another crucial aspect to the
scientific appropriateness of impact evaluation models is checking
beforehand the availability of baseline data, as well as, the (economic,
temporal, ethical) feasibility of measuring new data or collecting new
information throughout the monitoring process to get down the road.

2) Impact evaluation should be practical and straightforward but fulfil


technical requirements

Impact evaluation has to be practical and straightforward, including when


planned by scientists and conducted by experts. This implies that many barriers
should be overcome in communicating (and making aware of) the final aim of the
monitoring activity, to assure it is successful and well conducted.

41
SMART Specific, Measurable, Attributable, Realistic, and Timely or Time-bound, see Chapter 3

61
Since every NBS project is unique, measuring of impact/outcome needs to be
adjusted to that specific project and context. Although no universal framework
can be proposed, some basic requirements for a successful monitoring activity
are listed below.

• A high level, cooperative dialogue among practitioners, local or regional


authorities, stakeholders and scientists should occur from the beginning of
developing the monitoring and impact evaluation plans (see point 5) on
transdisciplinarity)
This will help practitioners, local or regional authorities and stakeholders to
be more aware about the critical aspects of a scientifically robust
assessment, as well as help scientists to focus more on the challenges that
really need to be tackled by the NBS intervention.

• Definition of the scope in which effects of the intervention are expected

• Definition of the site of investigation and/or target groups


The site of investigation can be the NBS site, its neighbourhood, its district,
the whole city or region. The target group is located within this spatial limit
and it should be as statistically representative as possible (see Chapter 3 and
Chapter 7).

• Choice of a control area/group (when applicable)


In many cases outside factors may influence outcome of the NBS
intervention. In order to validate the monitoring results and correlate them
with the NBS intervention realized, a parallel, twin, monitoring activity should
be performed elsewhere, by identifying the so-called “control area/group”. It
should be as identical as possible to the actual treated area/group. This
usually means that it should be located in the same
neighbourhood/district/city/region (depending on the scale at which effects
are expected, by scaling a level up the spatial scale) in order to take local
conditions (e.g., climatic conditions or cultural ones) into account. For
instance: if NBS effects are expected at the district level, the control
area/group should be chosen within the same city or region but in a different
district.

• Choice of a reliable and feasible frequency of data collection


Reliable frequency of the data collection should ensure the impact evaluation
on a temporal scale, which is adapted to the type of intervention and/or of
the challenge to be faced. However, data collection frequency should be also
feasible (see Figure 2-5), since regional authorities, municipalities or
stakeholders generally have limited budget/persons to do this.

3) Impact evaluations should clearly state and use reference conditions


and baseline assessment

62
Baseline data are important for measuring pre-intervention outcomes (reference
conditions) that are used later in the assessment process for the before-and-after
comparison. Chapter 7 of this handbook discusses how baseline data are
established and used operationally. In this section we list the following key points:

• Ensure that the method for establishing baseline data is repeatable

• Differentiate between process and outcome

• Chose standardized ways of assessing certain outcomes to allow for the


accumulation of evidence and comparability; striking a balance between
common indicators and highly specific ones;

• Assure clear link between challenges addressed and indicators selected

• Establish baseline and control area/group or reference values for


comparison in order to determine change(s) attributable to NBS
implementation

4) Impact evaluation should align with policy principles and reporting


obligations.

The expected outcomes based on objectives of an NBS intervention are important


for the impact evaluation. However, it is also important to identify and include
unexpected outcomes. Considering the time-frame of the project and the time
necessary for outcomes to be ‘visible’, some impacts may occur more quickly
than others.

In that sense, short-term immediately visible improvements are initial outcomes


that can be assessed immediately after the intervention (green quality, aesthetic,
amenities, etc.). Intermediate outcomes are assessable after some period of time
during the project (use and function of NBS, individual status and perception,
social environment) while long-term health outcomes (mortality rates, life
expectancy, cardiovascular disease, obesity, etc.) are often difficult to assess;
either because there is no long-term monitoring institutionalized, but also
because these outcomes are influenced by many interweaving factors. Moreover,
achieved positive impacts might change over time (depending on management,
succession, changing climate, etc.).

To assure relevance for policy-makers, it is also important to seek alignment with


key policy objectives. This can be done through a strategic review of policy
alignment between local/regional/national strategic objectives and potential NBS
benefits. The desired impact from the NBS implementation process can then feed
into the local administration, urban or regional policies (e.g., green roofs
mitigation and adaptation measure).

This should also provide connection to the local, national and EU-based policies
and requirements. For example, NATURA 2000 may require from all member
states to use certain indicators in the assessment of their natural areas. Similarly,

63
Floods Directive will specify those indicators that are related to flood risk
assessment. Water Framework Directive demands certain water quality
standards and indicators. Similarly, the LIFE programme 42, the EU’s funding
instrument for the environment and climate action, has developed a KPI
framework that can be seen as embedding element for measuring the impact of
a NBS. However, indicators in this Handbook (Chapter 4) are based on H2020
Projects involving EU and non-EU cities and regions and are thus applicable
globally.

5) Impact evaluation should be based on a transdisciplinary 43 approach.

Impact evaluation of NBS interventions relates to a whole range of different


societal challenges. It is unlikely that the knowledge required for such broad
evaluation sits with a single individual. As such, monitoring and evaluation teams
should engage societal actors and experts from across relevant disciplines in a
transdisciplinary approach. A transdisciplinary approach enables combining
knowledge from societal actors with knowledge and methods from different
disciplines (e.g., engineering, public health, social sciences, etc.) (Schneider et
al., 2019). To achieve transdisciplinarity, monitoring and evaluation plans should
be co-produced in collaborative actions to achieve the best balance between local
needs, values and knowledge, and scientific interdisciplinary knowledge and
requirements. Local authorities and practitioners, who are aware of real
conditions as well as administrative and technical barriers, should drive
collaborative actions. However, they should also involve additional expertise, for
example from the civic sector (to identify local needs and raise the awareness
about the benefits related to NBS), industry (to contribute to feasibility), and
scientists.

The co-production process should start with identifying a joint vision (Theory of
Change, Figure 2-2) and establishing desired outcomes collaboratively from the
beginning. By approaching co-production this way, it will be easier to relate
outcomes to the planned NBS, to expected results, and to the indicators that will
be used to measure the expected impact. Support from the local community is
crucial as this not only to improves the quality of information and trust in the
results of the impact evaluation itself, but also raises awareness and increases
sense of stewardship and caring. Likewise, partnerships and collaborations
among actors that are normally not in contact with each other can be generated.
Allowing different partners to get involved in participatory decision-making will
generate a sense of ownership of the solutions to be implemented (see also
Mahmoud and Morello, 2021).Their involvement will bring diverse perspectives in
defining outcomes, selecting indicators, collecting and analysing data.

Support from the scientific community or other experts is desirable when deciding
what methods or research designs will be considered credible for the impact
evaluation. This handbook is already driven by scientific principles and should

42
The LIFE Programme
43
Transdisciplinarity – problem-driven, cross-disciplinary, cooperative approach including scientists,
practitioners, stakeholders.

64
facilitate selection of suitable monitoring tools and protocols that can be adapted
to the local needs.

In that sense, it would be desirable that local administrations and practitioners in


collaboration with stakeholders and scientists interested in the implementation
and monitoring of a NBS:

• Tailor the monitoring protocols, while preserving the scientific robustness;

• Choose the needed experimental setup according to the required


resolution and disciplines; and,

• Follow up regarding the process during short and long-terms


implementation processes.

2.4 Capitalising on existing experiences and remaining critical


concerns

Impact evaluation of NBS interventions requires joint effort of different actors to


be able to assess wide range of outcomes and identify trade-offs before, during
and after the NBS implementation. A high-quality impact evaluation depends on
skills of team members conducting the study. However, even with a skilled team,
evaluation processes may face different challenges. In the following sections, we
describe challenges and gaps from H2020 projects and conclude with key
messages based on existing experiences from these projects.

2.4.1 Challenges and gaps in current monitoring and evaluation efforts

Impact evaluation is related to the interpretation of indicators selected to assess


NBS performance and effectiveness in addressing challenges and fulfilling
objectives. A number of common challenges and gaps in monitoring and
evaluation efforts are emerging from the existing NBS projects. These challenges
are analysed from four perspectives: practitioner, scientific, citizen/user and
private sector.

From a practitioner perspective main challenges are identified from project


work with stakeholders in cities and regions. They include a lack of expertise
in evaluation and data collection, in the critical selection of indicators that
address the predefined impacts; short time frames; dispersed and siloed data
within different agencies; lack of implementation monitoring vs. performance
monitoring (which could lead to the missing of important data afterwards,
such as for the accounting of the cost-benefit and cost-effectiveness); etc.
Problems of dispersed and siloed data can partly be solved with
transdisciplinary approach, which enables the effective gathering of data from
many different disciplines (health, air quality, biodiversity, water
management, economics, etc.) and effective communication with those who
hold those data.

65
The use of indicators themselves has following practical issues:

• Indicators exist but it is difficult to use them due to the lack of


understanding (e.g., understanding the logic behind the models), data
unavailability, data not available for use at fine scale (e.g., detailed
census data may be available at household level but cannot be released),
etc.

• Lack of resources, lack of ownership, lack of requirement from funders,


lack of interest once NBS has been installed, lack of expertise, change in
personnel

• Issues related to the complexity of cities and regions, as a system of


systems with several layers of networks constantly interacting with each
other, which makes it difficult to identify causal chains (especially when
people and their behaviour are the target of interest)

• The multiplicity of decision-making contexts and processes cannot be


captured by a universal and versatile set of indicators: each decision
requires the selection of ad-hoc indicators from among an extended set.
Formalisation of all those decisions is not always fully understood by the
different stakeholders who may expect easy ready-to–use methods
working in any conditions.

• Feasibility based on the available expertise (e.g., biomonitoring).

From a scientific perspective, (see section 2.3.2) the main gaps in the
monitoring process are:

• Lack of differentiation between the process and outcome, the gaps in the
monitoring methodology and implementation stages (micro-, meso-,
macro-, etc. scales of interventions) and longer-time frame of effects
measurement.

• Lack of longer-term evaluations to assess effects over time and


guaranteeing continuity of monitoring measurements: often models of
monitoring impacts lack the continuity of measurement from the pre-
greening to the long-term effects in the post-greening phase, they are
also influenced by the complexity and feasibility of the monitoring itself.
The ideal impact monitoring methodologies are the ones with the
minimum specialised equipment and time efforts, or relying on ready-to-
run and consolidated data acquisition protocols, possibly managed by the
public authority. Involving citizens and local stakeholders in the co-
monitoring of NBS interventions, often requires simplification, which is
challenging for some complex impacts.

• Difficulties in communicating to non-scientific partners in a less -technical


language. Engaging stakeholders in the process of data collection and
monitoring is challenging. However, scientists should translate indicators
to be simple and capable of immediate representation, easy to understand
and, connected to people‘s priority interests and concerns.

66
• Ability to express levels of uncertainty associated with evaluation
outcomes. Decision-makers want to know what is the relative level of
certainty or uncertainty associated with evaluation work. For example,
speaking in practical terms, if the likely chance of an NBS achieving its
intended impact is 80% then decision-makers may be very willing to up-
scale such an NBS intervention elsewhere, as opposed to their willingness
to upscale if the likelihood of achieving the desired impact is only 20%.

• Indicators exist but they may not be relevant to the studied NBS in a
place-based context. The way indicators are assessed (quantitative,
qualitative, traceability/justification of hypothesis) is essential.

• Any set of indicators will always remain contextual and correspond to the
knowledge level at a given moment: it is therefore interesting to provide
lists of indicators but also methodologies to build new ones in a dynamic
way if needed.

• Measurability of intangible impacts (e.g., aesthetic enjoyment) and


spillovers (impact of NBS intervention may spread beyond the treated
area or group) as well as accounting for trade-offs is challenging,
particularly because of the diverse perspectives of stakeholder valuing
NBS, the multiple time scales of assessment and influence of other
programs and factors.

• The assessment of NBS effectiveness or impacts is a multi-scale and


multi-temporal problem. Indicators for urban scales and issues may not
be relevant for wider scale such as catchment basin scale for example
when dealing with flood risk reduction.

• Indicators related to NBS effectiveness require the use of multi-


disciplinary approaches able to combine physical, environmental, social,
human and economic features. New paradigms are needed to integrate
this different kind of knowledge and related methods.

From citizens/users perspective: experience with citizen monitoring is limited


and collected data about the impacts of NBS is often not presented in a user-
friendly format and/or made available to the public. Need for scientific and
intercultural translation, lack of appropriation and adequate tools for co-
diagnostic, co-evaluation and co-monitoring that involve citizens as active actors
in the evaluation processes. Adoption of tools that include: the perception of
citizens, the translation and adaptation of content, the validation of monitoring
results by citizens. To consider people's voices, is to recognize the plurality and
open paths for effective co-production of knowledge, see section 2.3.2.

From a private sector perspective: in some cases, NBS are elaborated in


collaboration with industries and partners from the private sector. This is
particularly true when the NBS implementation includes regeneration of
previously productive sites and/or includes the implementation of innovation
technologies. In all these cases, to have valuable inputs, beyond the non-
monetisable benefits, is a real challenge.

67
In addition to the four perspectives, we identify three types of issues in NBS
implementation of monitoring and evaluation plans: technical, physical and
social. Some NBS which have been selected through the previous steps of building
a theory of change and which encompass an evaluation model (e.g., SMART)
have encountered a variety of hindrances in their actual implementation contexts,
such as:

• Technical issues: some NBS in place require a specific sophisticated technical


knowledge that is not necessarily available in project competences.

• Physical issues: some NBS in place have shown physical constraints or


drawbacks that might obstruct the implementation in reality or induce
unexpected side effects (e.g., a riparian forest causing woody debris and
bridges’ section reduction or even closure, see NAIAD project, La Brague
demonstration site).

• Social issues: a social acceptance factor towards implementation is


needed for any NBS impact model evaluation to measure an increase in
openness, awareness, citizen engagement and to assess management
efficiency, accountability, sharing, transparency, and communication.
That is why a transdisciplinary approach is needed in order to facilitate
the co-production of monitoring and evaluation plans with stakeholders.

In these cases, where the foreseen monitoring and evaluation plans cannot be
implemented, mitigation measures have to be applied.

2.4.2 Key messages from existing projects

NBS performance and impact evaluations should provide answers to policy


questions that affect people’s daily lives. In H2020 projects questions such as
‘Does an NBS intervention influence air quality, enable climate adaptation,
regulate microclimate, increase biodiversity or contribute to social cohesion and
well-being?’ are related to societal challenges. Key messages from these projects
are listed below.

Three core elements of well-designed NBS performance and impact evaluation are:

1. A concrete assessment question related to an outcome of interest


developed in a theory of change that can be answered with the impact
evaluation.

2. A robust methodology that balances understanding of the complexity of


diverse NBS outcomes, as well as trade-offs, with feasibility in relation to
the specific socio-economic context and available resources.

3. A well-formed evaluation team that functions as a transdisciplinary


partnership between different sectors (public, private, civil society) and
various knowledge disciplines depending on the type of NBS and outcomes
of interest.

68
It is important to have a practical focus and adapt these very general steps and
principles to local context and develop tailor-made monitoring and evaluation
plans. Moreover, don’t be afraid to start small and begin with evaluation
indicators that are more manageable and understandable. This can represent a
good foundation for the development of a transdisciplinary evaluation plan.

When developing such bespoke plans, although local practitioners and the local
population are crucial for plan development, it is also necessary to engage experts
from different disciplines to ensure that various benefits and co-benefits as well-
as unintended negative effects of NBS interventions are assessed and evaluated.
Although impact evaluations are complex processes with dynamic parts, they are
a worthwhile investment and collaboration can be the most effective way to
maximise the return on this investment.

Participants in the NBS impact evaluation should be included in the dissemination


efforts. Since they have invested their time and energy in planning and
implementing monitoring and evaluation plans, it is essential to ensure that they
have access to and remain informed about the evaluation results. This small effort
can contribute to their continued interest and willingness to participate in future
NBS evaluations.

On the following pages and between chapters there are different case studies
illustrating main characteristics and challenges of monitoring and evaluation
plans from different H2020 projects. Chapter 3 explains step-by-step the process
of development of monitoring and evaluation plans, which complements the
general overview provided in this chapter.

2.5 References

Baldacchini, C., Sgrigna, G., Clarke, W., Tallis, M., and Calfapietra, C., 'An ultra-spatially resolved method to
quali-quantitative monitor particulate matter in urban environment', Environmental Science and
Pollution Research, Vol. 26, 2019, pp. 18719–18729.

Bamberger, M., ‘Introduction to Mixed Methods in Impact Evaluation’, Impact Evaluation Notes, No 3, 2012.
Available from: https://www.interaction.org/wp-content/uploads/2019/03/Mixed-Methods-in-Impact-
Evaluation-English.pdf

CLEVER Cities project, D4.3 Monitoring strategy in the FR interventions, 2020. Available from:
https://clevercities.eu/fileadmin/user_upload/Resources/CLEVER_D4.3_Monitoring_Strategy_in_the
_FR_interventions_vF2.pdf

Dick, J., Carruthers-Jones, J., Carver, S., Dobel, A.J., and Miller, J.D., 'How are nature-based solutions
contributing to priority societal challenges surrounding human well-being in the United Kingdom: a
systematic map', Environmental Evidence, Vol. 9, 2020, pp. 1–21.

Dick, J., Miller, J.D., Carruthers-Jones, J., Dobel, A.J., Carver, S., Garbutt, A., Hester, A., Hails, R., Magreehan,
V., and Quinn, M., 'How are nature based solutions contributing to priority societal challenges
surrounding human well-being in the United Kingdom: A systematic map protocol', Environmental
Evidence, Vol. 8, 2019, pp. 1–11.

Funnell, S. and Rogers, P., Purposeful Program Theory: Effective Use of Theories of Change and Logic Models,
Jossey-Bass/Wiley, San Francisco, 2011.

Gertler, P.J., Martinez, S., Premand, P., Rawlings, L.B., and Vermeersch, C.M., Impact evaluation in Practice,
Second Edition, Inter-American Development Bank and World Bank, Washington, DC, 2016.
Available from: https://www.worldbank.org/en/programs/sief-trust-fund/publication/impact-
evaluation-in-practice

69
Harrison, P.A., Dunford, R., Barton, D.N., Kelemen, E., Martín-López, B., Norton, L., Termansen, M., Saaikoski,
H., Hendriks, K., Gómez-Baggethun, E., Czúcz, B., García-Llorente, M., Howard, D., Jacobs, S.,
Karlsen, M., Kopperoinen, L., Madsen, A., Rusch, G., van Eupen, M., Verweij, P., Smith, R.,
Tuomasjukka, D., and Zulian, G., ‘Selecting methods for ecosystem service assessment: A decision
tree approach’, Ecosystem Services, Vol. 29, 2018, pp. 481–498.

Langemeyer, J., Wedgwood, D., McPhearson, T., Baró, F., Madsen, A.L. and Barton, D.N., 'Creating urban
green infrastructure where it is needed – A spatial ecosystem service-based decision analysis of green
roofs in Barcelona', Science of The Total Environment, Vol. 707, 2020, 135487.

Mahmoud, I. and Morello, E., 'Co-creation Pathway for Urban Nature-Based Solutions: Testing a Shared-
Governance Approach in Three Cities and Nine Action Labs', Smart and Sustainable Planning for Cities
and Regions, Springer International Publishing, 2021, pp. 259–276.

Morra Imas, L.G. and Rist, R., The Road to Results: Designing and Conducting Effective Development
Evaluations, World Bank, 2009.

Morton, M.H., Applicability of Impact Evaluation to Cohesion Policy, Report Working Paper, 2009. Available
from: https://ec.europa.eu/regional_policy/archive/policy/future/pdf/4_morton_final-formatted.pdf

Pintér, L., Hardi, P., Martinuzzi, A., and Hall, J., ‘Bellagio STAMP: Principles for sustainability assessment and
measurement’, Ecological Indicators, Vol. 17, 2012, pp. 20-28.

ProGIreg, Methodology on spatial analysis in front-runner and follower cities, 2018. Available from:
https://progireg.eu/resources/planning-implementing-nbs/

Raymond, C.M., Berry, P., Breil, M., Nita, M.R., Kabisch, N., de Bel, M., Enzi, V., Frantzeskaki, N., Geneletti,
D., Cardinaletti, M., Lovinger, L., Basnou, C., Monteiro, A., Robrecht, H., Sgrigna, G., Munari, L. and
Calfapietra, C., An Impact Evaluation Framework to Support Planning and Evaluation of Nature-based
Solutions Projects, An EKLIPSE Expert Working Group report, Centre for Ecology and Hydrology,
Wallingford, 2017.

Rogers, P.J., ‘Matching Impact Evaluation Design to the Nature of the Intervention and the Purpose of the
Evaluation’, Journal of Development Effectiveness, Vol. 1, No 3, 2009, pp. 217- 226.

Schneider, F., Giger, M., Harari, N., Moser, S., Oberlack, C., Providoli, I., Schmid, L., Tribaldos, T. and
Zimmermann, A., ‘Transdisciplinary co-production of knowledge and sustainability transformations:
Three generic mechanisms of impact generation’, Environmental Science and Policy, Vol. 102, 2019,
pp. 26-35.

Tacnet, J.-M., Piton, G., Favier, P., Pengal, P., Curt, C., Yordanova, R., Van Cauwenbergh, N., Giordano, R.,
Natural Based Solutions choice and effectiveness assessment: Integrative modelling and decision-
aiding framework, Editions Quae, Versailles, 2021 (submitted).

van Oudenhoven, A.P., Schröter, M., Drakou, E.G., Geijzendorffer, I.R., Jacobs, S., van Bodegom, P.M.,
Chazee, L., Czúcz, B., Grunewald, K., Lillebø, A.I., Mononen, L., Nogueira, A.J.A., Pacheco-Romero,
M., Perennou, C., Remme, R.P., Rova, S., Sybre, R.-U., Tratalos, J.A., Vallejos, M., and Albert, C.,
‘Key criteria for developing ecosystem service indicators to inform decision making’, Ecological
Indicators, Vol. 95, No 1, 2018, pp. 417-426.

White, S. and Pettit, J., ‘Participatory Methods and the Measurement of Wellbeing’, Participatory Learning and
Action, Vol. 50, 2004, pp. 88-96.

70
Drawing on knowlegde from projects
funded by the European Union

CONNECTING NATURE
Bringing cities to life, bringing life into cities
Genk (BE) Glasgow (GB) Poznań (PL) A Coruña (ES) Burgas (BG)
Ioannina (GR) Málaga (ES) Nicosia (CY) Pavlos Melas (GR) Sarajevo (BA)

Brings in actions to feed the initiation and expansion of economic and social enterprises in
production and large-scale implementation of NBS in urban settings to measure the impact
of these initiatives on climate change adaptation, health and well-being, social cohesion and
sustainable economic development.

Image: Kindergarten, Poznań - Photo © Piotr Bedliński, City of Poznań


SCOPE NBS at city scale

Approach to Impact Assessment Main Challenges addressed

The project developed indicators working


co-creatively with our Front-runner cities. The
1. Climate Resilience
goals behind both past and planned NBS pro-
jects, and the associated benefits and co-bene- 2. Water Management
fits delivered were explored. These criteria were
used to filter a comprehensive list of indicators 3. Natural and Climate Hazards
to identify which were recommended ‚core‘ (i.e.,
had key relevance for all cities across all NBS), 4. Green Space Management
and which were additional ‚feature‘ (i.e., had
potential relevance, but were more specific to in- 5. Biodiversity
dividual NBS). The results were sense-checked
with the city teams, and then different options 6. Air Quality
were developed for implementing each indicator
dependent upon the expertise and capacity wit- 7. Place Regeneration
hin each city. The indicators were implemented
in the cities, to ensure that a diverse evidence 8. Knowledge and Social Capacity Building
base of benefits is available to unlock broader
funding and secure political buy-in. 9. Participatory Planning and Governance

10. Social Justice and Social Cohesion


Involved Stakeholders and roles
11. Health and Wellbeing
The stakeholders involved in the Connecting Nature
NBS impact assessment process were the following: 12. New Economic Opportunities & Green Jobs

• City planners and officers: they were in charge


of detailing the NBS, as well as participating in
the adaptation of the indicators to their cities. Lessons learned
They were also the link to local organizations
and communities for data collection. A key challenge has been the diversity of expertise
• Science-practice partners: academics belon- within the city teams in relation to evaluation indi-
ging to universities, who guided cities through cators. Some cities had little evaluation expertise
the steps of the NBS impact assessment pro- in the area of nature-based solutions, while for ot-
cess. The collaboration was established through hers expertise in one area such as environmental
periodic meetings where the evaluation and indicators existed, but it was either housed in a
monitoring plans for each city were developed. different department or was provided by a univer-
• Consultant companies and SMEs: they had the sity partner. One of the main lessons we all learned
function of interface with the cities to locate the was how working co-productively with local teams,
stored knowledge of the NBS, and co-produce through regular meetings and mutual agreements
NBS catalogues and monitoring tools. for each of the impact assessment steps. In this
way, training activities were carried out for cities
in evaluation and monitoring competencies, but
in addition, the importance of the impact assess-
ment process was also explained and emphasized.
As the local teams became empowered to design
their NBS monitoring and evaluation plans, they
Municipal Administrations (FR/FL) increased their efforts aimed at understanding the
real impact of their urban interventions.
Planning experts

Scientists / Academia

Schools and kindergartens


Learn more
Consultant Companies and SMEs https://connectingnature.eu/

The Connecting Nature project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 730222
Grow Green
Manchester (GB) Valencia (ES) Wroclaw (PL)
Brest (FR) Modena (IT) Zadar (HR) Wuhan (CN)

The project aims to accelerate the delivery of NBS strategies across Cities. By investing
in NBS pilot projects in Manchester, Valencia and Wroclaw that deliver quantified im-
provements in climate and water resilience, social, environmental and economic perfor-
mance, the project will develop a robust evidence base and a replicable approach that
will enable this acceleration across Europe and the rest of the world.

Drawing on knowlegde from projects


funded by the European Union
Image: West Gorton Community Park, Manchester - Photo © Manchester City Council
SCOPE A partnership for greener cities to increase
liveability, sustainability and business opportunities

Approach to Impact Assessment Main Challenges addressed

The impact assessment will be undertaken at


two different levels. At a city level the impact
1. Climate Resilience
of each pilot project will be evaluated in terms
of evidence-based outcomes, key messages 2. Water Management
and lessons learned.
3. Natural and Climate Hazards
A thematic evaluation of specific NBS inter-
ventions will also be undertaken based on the 4. Green Space Management
Eklipse framework challenges of climate resi-
lience, water management, green space ma- 5. Biodiversity
nagement, bio diversity, air quality, social jus-
tice and social cohesion, health and wellbeing, 6. Air Quality
economic opportunities and green jobs.
7. Place Regeneration

8. Knowledge and Social Capacity Building

9. Participatory Planning and Governance

Involved Stakeholders and roles 10. Social Justice and Social Cohesion

The stakeholders involved for the monitoring 11. Health and Wellbeing
process provides a rich co monitoring opportuni-
ties: Civil society – citizens and representatives 12. New Economic Opportunities & Green Jobs
of active associations, private sector, Academia
policy makers and public sector/associated ser-
vice stakeholders. Nevertheless the degree of
engagement and interaction of each type of sta- Lessons learned
keholders depends on the cities’ requirements
and culture about participation.
The EKLIPSE framework is the basis for the
KPIs identification but to assure the alignment
Municipal Administrations of the monitoring strategy with the expected
outcomes, local stakeholders must be integra-
ted in the process since the beginning.
Regional/national statistics authority
Climate related variables has specific conditio-
Citizen ning for monitoring due to scale (space and time
domains) that must be considered to plan the
monitoring strategy. For some KPIs or variables
Planning experts modelling could offer a rich information to fill
some monitoring GAPs or to avoid uncertainty.
Scientists / Academia

NGOs

Schools and kindergartens


Learn more
www.growgreenproject.eu/

The GROWGREEN project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 730283
UNaLab
Urban Nature Labs
Eindhoven (NL) Tampere (FI) Genoa (IT) Stavanger (NO) Prague (CZ)
Castellón (ES) Cannes (FR) Başakşehir (TR) Hong Kong (CN) Buenos Aires (BR)

UNaLab is generating evidence of the benefits, cost-effectiveness, economic viability and


replicability of NBS targeting climate change mitigation and adaptation, and sustainable
water management. UNaLab activities promote smart, inclusive, resilient and sustainable
urban communities through stakeholder co-creation of Urban Living Labs (ULLs) and local
NBS demonstrations, and co-evaluation of NBS impact. Collaborative knowledge production
among the network of UNaLab partner cities yields project results that reflect diverse urban
socio-economic realities, along with differences in the size and density of urban populati-
ons, local ecosystem characteristics and climate conditions. UNaLab project outcomes that
support further replication and up-scaling of NBS include an ULL model, ICT tools for NBS
co-creation and co-monitoring, applicable business and financing models, and guidance on
governance-related structures and processes to support NBS uptake.

Drawing on knowlegde from projects


funded by the European Union
Image: Vuores Central Park, Tampere – Photo © City of Tampere
SCOPE NBS for climate- and water-resilient urban areas

Approach to Impact Assessment Main Challenges addressed

UNaLab uses a highly participatory approach


to produce evidence of NBS impact, including
1. Climate Resilience
co-creation, co-development, and co-monitoring
activities. The impact assessment of the NBS 2. Water Management
in the UNaLab front-runner cities first involved
iterative co-definition of Key Performance Indi- 3. Natural and Climate Hazards
cators (KPIs) and Key Impact Indicators (KIIs)
with a wide range of stakeholders. The UNaLab 4. Green Space Management
front-runner cities then iteratively co-developed
robust monitoring and evaluation strategies to- 5. Biodiversity
gether with project partners and other technical
experts to thoroughly assess NBS performance 6. Air Quality
and impacts in a cost-effective way. The ICT
platform and NBS monitoring and evaluation 7. Place Regeneration
tools developed by UNaLab project partners
support long-term NBS evaluation by enabling 8. Knowledge and Social Capacity Building
the automated collection of monitoring data for
NBS impact assessment from IoT sensors whe- 9. Participatory Planning and Governance
rever possible, while allowing manual entry of
data as needed. 10. Social Justice and Social Cohesion

11. Health and Wellbeing


Involved Stakeholders and roles 12. New Economic Opportunities & Green Jobs

Stakeholders played a critical role in developing


and shaping the UNaLab impact assessment fra-
mework. Local stakeholders, project partners and Lessons learned
external experts in each UNaLab front-runner
city co-identified the Key Performance Indicators Co-development of the monitoring strategy relies
(KPIs) and Key Impact Indicators (KIIs) based on on a diversity of participants, in terms of cultural
the local challenges and SMART criteria in a series and educational background and needs, which
of group sessions. These sessions shaped the com- requires on-going communication to maintain
mon understanding of challenges and their relati- active participant engagement. Steadfast stake-
ve importance, as well as the expected outcomes holder engagement proved to be essential for
of planned NBS actions in each city. After several identifying the local challenges and monitoring
iterative cycles of indicator selection, project part- and evaluation needs. Challenges in the defini-
ners co-developed a final set of common indicators tion of performance and impact indicators and
for NBS monitoring and evaluation. Stakeholders monitoring needs were addressed through enga-
further participated in the selection of appropriate gement of a wide range of experts during NBS
monitoring protocols and the development of local assessment planning. Obstacles encountered du-
co-management activities such as the engagement ring the NBS implementation process influenced
of students in NBS monitoring and the establish- the final set of impact assessment indicators for
ment of local Communities of Practice. a few selected NBS.

Municipal Administrations

Citizen

Planning experts

Scientists / Academia
Learn more
www.unalab.eu
NGOs

The UNaLab project has received funding from the European Union’s Horizon 2020
research and innovation programme under Grant Agreement No. 730052
URBAN GreenUP
New strategy for Renaturing Cities
through Nature-based Solutions
Valladolid (ES) Liverpool (GB) Ìzmir (TR) Ludwigsburg (DE) Mantova (IT)
Medellin (CO) Chengdu (CN) Binh Dinh - Quy Nhon (VN)

URBAN GreenUP project wants to develop a new concept, “Renaturing Urban Plans (RUPs)”, which
include actions focused on mitigating the effects and risks of climate change and improving the air
quality and water management of cities. The urban renaturing methodology developed by URBAN
GreenUP is demonstrated in three front-runner cities, Liverpool (The UK), Izmir (Turkey) and Valla-
dolid (Spain). Based on their experience, five follower cities will set up their own Renaturing Urban
Plans to replicate the URBAN GreenUP strategy and act as ambassadors for a broader group of cities
with a high replication potential. The main objectives of URBAN greenUP are to (1) develop and
demonstrate a fully replicable renaturing methodology to support the development of Renaturing
Urban Plans aimed at climate change mitigation and efficient water management; (2) involve citi-
zens, local authorities and stakeholders in the co-design of their city renaturing plans; (3) identify
innovative business plans to replicate the model in other cities all around the world; (4) foster the
creation of a global NBS market and support EU international cooperation.

Drawing on knowlegde from projects


funded by the European Union
Image: Green cover Plaza España, Valladolid - Photo © Valladolid City Council
SCOPE Nature Based Solutions for renaturing cities

Approach to Impact Assessment Main Challenges addressed

The key aim is to quantify the impacts of NBS in the ci-


ties to enhance the quality of life of the citizen through
1. Climate Resilience
measuring multiple axes, following the significant
principles of effectiveness, repeatable and reasona- 2. Water Management
ble cost. Each city partner must focus in their precise
goals and aim for a monitoring program that tackles 3. Natural and Climate Hazards
the main issues and challenges that each city is facing.
Key effort during the monitoring program is to learn 4. Green Space Management
lessons from the process, draw data through different
sources and cities and derive global conclusions that 5. Biodiversity
will serve the main objectives of the Project as im-
proving citizen well-being and palliate climate change 6. Air Quality
effects in cities. The monitoring program will serve as
evaluation tool for not only the NBS per se but for the 7. Place Regeneration
Project itself for many reasons. In one hand, KPIs will
provide information regarding NBS but also will collect 8. Knowledge and Social Capacity Building
data that can be used to calculate city-wide indicators
that apart from serve as NBS indicator can be used to 9. Participatory Planning and Governance
determine further evaluation and global conclusions.
10. Social Justice and Social Cohesion

Involved Stakeholders and roles 11. Health and Wellbeing

The monitoring description and the description of 12. New Economic Opportunities & Green Jobs
the KPIs can be utilized by:
• Demo Cities and municipal administrations,
enabling them to develop strategies based on
the progress of the NBS. Lessons learned
• City residents and non –profit citizen organi-
zations enabling them to understand the de- All issues encountered during the monitoring program
velopment and the baseline of the city. are shared and dealt with by all the partners involved
• Follower cities, in order to learn from the use in order to find the best possible solution.
and application of the NBS and the improve- • Storage requirements for some KPI data and
ment on the cities. who is in charge. In general, cities are in charge
• Other professionals of urban planning, geogra- of the data storage.
phers, architects and landscape professionals. • Coordination between who is in charge of
what: obtain raw data, calculation KPI, output
data owner. Partners led by the Monitoring WP
accorded responsibilities, defining the roles of
Municipal Administrations (FR/FL) the different partners.
• Different timing between cities and implemen-
Regional/national statistics authority tations due the tendering processes. Internally
managed by the front-runner cities.
Citizen
As main lessons learned we can consider as following:
• The generation of participatory process between
Planning experts experts and partners
• The data provide useful knowledge for stakehol-
Scientists / Academia ders beyond the purpose of the Project
• The need for storage requirements for all the
NGOs data produced Role definition is required for the
performance of the monitoring process.
Green businesses
Learn more
www.urbangreenup.eu
Schools and kindergartens

The URBAN GreenUP project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 730426
3 APPROACHES TO MONITORING
AND EVALUATION STRATEGY DEVELOPMENT
Coordinating Lead authors
Dumitru, A., Garcia, I., Zorita, S., Tomé-Lourido, D.

Contributing authors
Cardinali, M., Feliu, E., Fermoso, J., Ferilli, G., Guidolotti, G., Hölscher, K., Lodder, M.,
Reichborn-Kjennerud, K., Rinta-Hiiro, V., Maia, S.

Summary

What is this chapter of the Handbook about?

In this chapter, we outline a step-by-step approach to developing and


implementing an impact assessment plan that covers all stages from planning
and implementing to achieving policy impact. Understanding the specific steps to
consider and follow when planning and implementing evaluation, will help
practitioners make appropriate on-the-ground decisions that fit to their local
context

We begin with introducing a structured reflection process that connects your


strategic objectives, with NBS actions and expected outcomes, through the
mapping of a theory of change, and the development of a logical chain of results
that differentiates between process characteristics and outcomes (Section 3.1).
We then delve into the steps involved in designing effective monitoring and
evaluation plans (Section 3.2). Next we outline the key features and conditions

79
needed for a successful process of co-production of monitoring and evaluation
plans, involving a diversity of stakeholders, from a quintuple helix perspective
(Section 3.3). Finally, we present three innovative tools oriented to enhancing
reflexivity in impact assessment and NBS design and implementation, more
generally; to support the development of tailored monitoring and evaluation plans
for local NBS; and to gather user data with the support of automatized procedures
and technological devices (Section 3.4). The chapter concludes by stressing the
role of robust monitoring and evaluation in evidence-based policy-making, the
creation of a culture of continuous evaluation, and in stakeholder and citizen
education (Section 3.5).

How do I use this chapter in my work with NBS?

You can use this chapter to develop your impact assessment strategy from the
beginning of your NBS planning process. The chapter also outlines how
monitoring and evaluation plans can feed into wider assessment, data collection,
and reporting efforts, with a long-term view.

When should I use this knowledge in my work with NBS?

Monitoring and evaluation is sometimes considered too late in the process of NBS
implementation that important opportunities are lost because of it. Therefore, we
recommend that you use this chapter at the beginning of your planning process:
it will enable you to have an overview of the steps you need to follow and thus
save time and resources by initiating certain actions and collaborations early in
the process. It might also be useful to review each step as you go through them,
to ensure that you have considered all relevant aspects in each stage.

How does this chapter link with the other parts of the handbook?

After the in-depth description of principles that should be followed in developing


robust impact assessment in chapter 2, this chapter describes the practical steps
in detail, and outlines how impact assessment can be done through adopting a
co-production approach. Specific indicators for each challenge category are then
described in chapter 4. Considerations regarding data are discussed in chapter 5.

3.1 Introduction: developing robust impact assessment plans

Robust impact assessment is a key aspect of the urban and regional regeneration
and resilience agenda in Europe. Nature-based solutions have emerged as a
promising and potentially effective type of interventions for a variety of
environmental, social and economic challenges. However, clear and sufficient

80
evidence on their different outcomes, the synergies and trade-offs between
these, and the processes and pathways through which outcomes are achieved is
still needed (Dumitru et al., 2020). Robust evaluation of nature-based solutions
(NBS) in different cities and regions will contribute to an evidence base that can
inform urban planning and interventions, investments and policy-making. In the
medium and long term, it can contribute to the creation of a culture of impact
assessment, as part of the design and implementation of nature-based and grey
solutions.
As participants in the large-scale EC H2020 NBS projects described throughout
this handbook, many cities and regions are defining local NBS monitoring and
assessment plans and facing numerous challenges. Robust monitoring and
evaluation plans provide important knowledge regarding the strengths and
weaknesses of nature-based interventions, and the degree of achievement of the
strategic objectives of the stakeholders involved. The effective development and
implementation of these plans requires a thoughtful, step-by-step approach and
active collaboration with local stakeholders. It is not a task that should be carried
out in isolation, and this chapter seeks to offer orientation by describing in detail
the step by step approach to monitoring and evaluation briefly outlined in Chapter
2, as well as outlining the key characteristics and stages involved in a co-
production approach to impact assessment design and implementation.
Effective monitoring and evaluation plans have been identified as a key enabler
for successful implementation of NBS (Ershad-Sarabi et al., 2019). In fact, when
impact assessment plans follow, and are aligned with, local spatial development
objectives, they support the transition to natured-based solutions design, by
providing the evidence base for projects, plans and policies (Geneletti et al.,
2016).
Collaborations between scientific experts, municipalities and other stakeholders
are particularly helpful in the development and implementation of such robust
impact assessment plans. Collaboration with local universities or urban
professionals with scientific knowledge and experience is very valuable, as
nature-based solutions have impacts across a wide range of contemporary
challenges, thus requiring a wide range of scientific expertise (Raymond et al.,
2017b). Successful co-creation experiences between researchers and policy
officers in the design, implementation and maintenance of nature-based solutions
leads to mutual learning and the establishment of relationships of trust
(Frantzeskaki and Kabisch, 2016), facilitating long-term collaboration.

3.2 A step by step approach to developing robust monitoring and


evaluation plans for NBS

A robust monitoring and evaluation strategy requires careful planning from the
beginning of the process of NBS design. By following a step-by-step approach,
adequate resources can be assigned. To make sure evaluation is both robust and
cost-effective. Teams in charge of developing and implementing a nature-based
solution can work through a series of six sequential steps, already briefly
summarized in Chapter 2. The process is not entirely linear, and feedback loops
between some of the steps exist, as described below. A synthesis of these six
steps and the relationships between them is presented in Figure 3.1, illustrating

81
how constructing a theory of change is an iterative process, and the feedback
loop between steps 2 (outlining the sequence of results) and step 3 (specifying
impact), which will feed into and help refine step 1 (the theory of change).

Figure 3-1. Summary of steps for developing impact monitoring and evaluation plans

STEP 1: Constructing a theory of change

The development of a theory of change enables planners and decision-makers to


establish a clear relationship between key local context challenges, strategic
objectives and the actions through which these will be reached, and fosters clear
identification and reflection on the linkages, or pathways, between them.
Developing a good theory of change takes time, but this effort will pay off in
subsequent stages of monitoring and evaluation planning, by saving considerable
time and money, through the anticipation and mitigation of errors. The following
stages can be identified when developing a theory of change:

1.a) Engage in structured reflection on key local context challenges and NBS
objectives
Structured reflection supports cities in establishing context-appropriate rationales
for NBS implementation and establishing impact assessment objectives (Dumitru
et al., 2021). Strategic objectives in a particular city or region are normally
implemented by establishing more specific, local goals, and by identifying local
challenges that call for specific policy interventions to achieve those goals.
Developing a theory of change entails making these relations explicit with some
degree of formalization, by providing answers to the following questions: which
local goals are targeted; what city or regional strategic objectives they address;
what nature-based solution/s and actions will address them; what, what specific

82
outcomes are expected at different stages of the change process and which
specific outputs will be sought to achieve those outcomes.
Strategic goals are normally defined in strategic policy documents and defined in
broad terms. Fitting or relating these to international targets such as the
Sustainable Development Goals (SDGs) of the United Nations 2030 Agenda for
Sustainable Development (2015) is helpful in adopting a bigger picture view of
strategic objectives that will be addressed, among other, by NBS interventions
and contributes to establishing connections between monitoring and evaluation
efforts that are already taking place in the city or region. It also provides
arguments to enhance collaborations between different stakeholders and acquire
necessary funds for monitoring and evaluation.
A clear relationship should be established between specific NBS outcomes and
the actions that need to be implemented at different stages, to produce those
outcomes. Specific outputs should be listed for each of these actions and
stakeholders should spend some time reflecting on potential interactions
between outcomes that might lead to both positive synergies and unwanted
trade-offs.

1.b) Involve the appropriate stakeholders and foster a sense of belonging to the
process
Each stakeholder might have a different vision of the objectives to be set, the
way to achieve them, or knowledge about the likelihood of different pathways
connecting interventions to outcomes. Stakeholders also bring informed
perspectives on local needs, as well as visions of the desired transformation and
the role of NBS in achieving it. These points of view are not exclusive but
complementary and will enrich the theory of change. An additional benefit of an
approach that involves stakeholders from the beginning is that it fosters active
engagement and a sense of belonging among stakeholders, as well as
relationships of trust and cooperation (see Section 3.3 for additional detail).
Local teams responsible for monitoring and evaluation will benefit from holding
regular meetings with stakeholders, in an iterative process. The vision of
decision-makers will likely be enriched by other stakeholders’ needs, desires,
expertise and feedback on what may or may not work, and on the outputs and
outcomes needed to achieve strategic goals and effectively address local
challenges.
The presence of technical staff or a group of monitoring experts is important
across the whole process of monitoring and evaluation, at varying intensities.
Experts might be specialists in different categories of impacts or challenge areas,
or in co-production activities, and they might also advise on the customization of
the impact assessment plan to the capacities and resources of the city. Many
times, local teams already have some technical expertise among their staff, which
may be complemented with external resources, such as collaborations with
scientists and universities. Experts’ contribution will be essential in later stages
of planning, when expertise on impact assessment methodologies and data
collection is needed.

83
STEP 2: Developing a results chain to outline the theory of change
Following the clarification of local challenges, key local goals, and NBS actions to
achieve them, stakeholders should explicitly identify assumptions regarding the
mechanisms by which NBS actions will lead to expected impacts. Explicitly
mapping the expected causal chain by which the implementation of the NBS will
achieve strategic objectives, is useful in anticipating what may be missing in the
design. Mapping causal pathways also allows for early detection of situations
where NBS might not deliver all the envisioned outcomes, and beginning to ask
the right questions about why that might be the case. Such a reflexive approach
also fosters experimentation with tweaking design or with additional measures to
improve NBS effectiveness over time.
When mapping causal pathways, the intermediary pathways through which an
NBS, an NBS feature or an NBS action might lead to the expected outputs and
outcomes should be clearly specified. Outcomes are the concrete results sought
through the implementation of an NBS (e.g., reduce air temperature or increase
mental health and wellbeing), while outputs are the visible part of NBS
interventions necessary to fulfil the outcomes (e.g., create an urban green park;
implement a participatory process of NBS design). The city has explicitly
established its assumptions when it has achieved clarity, and can specify what
actions will be carried out, what results are expected to be achieved through
them, and what they think are the mechanisms that explain why an action is
likely to lead to a particular outcome or result.
Imagine, for example, a neighbourhood who defines a series of strategic
objectives of improving levels of physical activity in youth, and decides to create
a neighbourhood park that would allow for people to be outdoors and exercise.
In some cases, the assumption is that having the park in place would create
recreational and exercise opportunities for youth, thus establishing a direct causal
pathway between the existence of the park and physical activities. However,
imagine now that the park is not accessible to a part of the neighbourhood
because it does not have sufficient access points, or that particular socio-
demographic groups such as cultural minorities or young women do not use the
park as they do not feel safe in it. We start to see that we might need to consider
additional pathways or conditions that lead to the expected outcome, such as
accessibility of the park or perceived safety, and include them in the assessment.
Furthermore, two types of impacts can be distinguished. “Intended” impacts are
the effects or changes that are not only desirable but are explicitly targeted
through the NBS implementation. “Unintended” impacts are the (usually)
negative, unforeseen results of NBS implementation. Also, each local team should
establish its theory of change based on knowledge of the local context, since
there are many factors that can influence the successful achievement of outputs
and outcomes. Sometimes there are interrelationships of "positive effects", also
called synergies (e.g., creating large tracts of urban green spaces favours
biodiversity but also offers spaces for physical activity), while in other cases,
there may be interrelationships of "negative effects" or trade-offs (e.g., creating
parks that improve the perceived quality of urban environments, which in turn
contributes to gentrification, and the exclusion of some groups).
Local teams should reflect upon and identify the possible intended and
unintended impacts, as well as synergies and trade-offs that may occur across

84
the causal pathway. This will be of great importance in assigning causality, as
described below.

STEP 3: Specifying the evaluation question(s)


The main reason for the development of robust NBS monitoring and evaluation
plans is to establish the direct effect that these interventions have on addressing
particular challenges and reaching certain objectives. As described in Chapter 2,
impact evaluation is about answering causal questions: To what extent is this
park contributing to reductions of obesity in a neighbourhood? To what extent is
this urban garden contributing to reductions of depression rates in this
neighbourhood, and through which mechanisms does it do so? Is it through
increased physical activity, through simple exposure to nature, or through the
fostering of increased contact and positive interactions between users? To what
extent is this intervention more effective (if at all), than no intervention (where
depression rates might improve anyway with the passing of time), or as
compared to alternative, non-NBS interventions? Making these questions specific
provides narrative context to the theory of change and orients the choice of
appropriate indicators.
It is also useful to identify other factors that might influence the same outcomes
in a given location and time period, as well as the relationship between NBS
actions and outcomes. Some of these factors will be beyond decision-makers’
control, but anticipating at least some of them will help with the correct
attribution of causality, or, said differently, with knowing which are directly
attributable to the NBS and which are not. Different options to correctly establish
causal relations between NBS actions and outcomes have been outlined in
Chapter 2.

STEP 4: Selecting indicators and data gathering methods - assessment


of performance and process
Adequate indicators should allow for the assessment of both performance and
process, and thus answer the following questions: does the NBS operate as
designed and are outcomes consistent with the planned theory of change?

4.a) Select appropriate indicators

Throughout this handbook, indicators associated with 12 societal challenge areas


(e.g., climate resilience, health and well-being, etc.) are presented. Each of these
indicators has been developed using SMART (Specific, Measurable, Attributable,
Realistic, and Time-bound) criteria, and each refers to the assessment of
particular outcomes. Process indicators are also included, which refer to the
characteristics of the NBS implementation process (e.g., number of stakeholders
involved in the initial NBS design stage). When indicators are selected to assess
one or several NBS projects, together they should form a coherent framework,
considering the synergies and trade-offs mapped in the theory of change. In some
cases, it is difficult to choose and measure all the desired outcomes and process
features outlined in the previous steps, due to constraints in financial, human and

85
time resources. Therefore, in collaboration with the stakeholders, indicators will
need to be ranked to establish priorities, to differentiate between those that are
critical to the assessment of key NBS expected outcomes (recommended, or core,
indicators) and those that might be desirable when additional resources and
stakeholder collaborations are available and possible (additional indicators).
For each of the 12 challenge areas selected, Chapter 4 presents a set of
recommended indicators, considered essential to mapping key outcomes of
different types of nature-based solutions, and a set of additional indicators that
might fit certain local contexts and types of nature-based solutions, but not
others. Aware of the fact that resources are always limited to some extent, the
list of core indicators has been kept to a minimum, while the list of additional
indicators include a wide range of outcomes, and scientifically valid methods for
their assessment. Local teams can start with the core indicators and progressively
expand it over time, in line with policy priorities and resources.
Local teams can graphically illustrate which indicators are chosen for each of the
important assumptions in their theory of change, through the use of causal maps,
as illustrated by an example from the Connecting Nature project, presented in
Figure 3-2.

Figure 3-2. Indicator causal map


(adapted from Dumitru et al., 2021; approach used in the H2020 Connecting Nature project)

4.b) Choose an appropriate impact evaluation method

Once the indicators have been selected, within a coherent framework, the next
phase will consist of identifying an appropriate method for each indicator. There
may be more than one measurement method for each indicator (e.g., physical
activity can be measured through a self-reported questionnaire, wearable devices
or through heat maps). For each of the indicators presented in this Handbook, at
least one measurement method is proposed. For those cases where end-users
have to make decisions between several options, and choose a method adapted
to their characteristics, the following three criteria outlined in Table 3-1 should
be considered.

86
Table 3-1. Factors influencing selection of NBS impact evaluation measurement methods

Data quality Involves the selection of standardized, scientifically-tested


measurement instruments. High data quality is critical to
enable drawing of valid conclusions, especially related to
causality.

Temporal adequacy Some NBS impacts will be registered shortly after NBS
implementation, while others will take time. For example,
reduction in the prevalence or incidence of different
illnesses might need a long time span of 5-10 years to be
registered. Frequency and temporal planning of
measurements should take these aspects into account.

Cost-benefit ratio Some methodologies provide highly detailed and accurate


data but are very costly. When a particular impact is
important for the city, or when over-time benefits are highly
proportional to costs, these should be considered. High-
quality, precise data pays off in the long term.

4.c) Identify and collect the data needed to assess selected indicators

After selecting appropriate indicators and methodologies, the next step is to


identify available data and decide in which cases new data should be collected.
In the previous chapter, the difference between baseline (prior to NBS
implementation) and outcome data (data subsequent to NBS implementation)
was explained (see also Figure 3.3). The absence of baseline data considerably
limits the possibility of attributing impacts to the implementation of the NBS.
Certain relationships may be observed, but it will be impossible to know for sure
whether they are due to the NBS, or whether they might be due to other co-
occurring phenomena.

Figure 3-3. Baseline vs Outcome data (adapted from Dumitru et al., 2021)

It is strongly recommended to either detect data sources for the baseline and
then collect outcome data, or, where data is not available, plan for baseline data
collection before NBS design and implementation takes place. Moreover, given
adequate resources, as well as the possibilities afforded by certain automatized

87
forms of data collection (such as wearable or remote sensors, smartphones, etc.)
data might be also collected at several times before, during and after NBS
implementation, thus allowing for higher precision and the detection of subtle
variations as a result of NBS implementation.
In some cases, data is already available through public, private or third sector
agencies at national or international levels. Thoroughly reviewing available data,
as well as attempting to connect data collection with existing and regular survey,
monitoring and reporting efforts at regional, national or international levels will
mean that monitoring and evaluation of NBS can become a regular practice and
be maintained and enriched over time.

4.d) Developing a local monitoring and data collection plan

The development of an effective local monitoring plan should consider a


structured sequence of actions (CLES, 2010; Compass, 2010; United Nations,
2009), that together form a coherent data collection plan, with specific
requirements regarding types of data, target populations and samples to be used,
specific data analysis techniques and provisions for the protection and storage of
data. Questions that the monitoring and data collection plan should answer are
shown in Table 3-2. First, stakeholders should be assigned different roles in the
monitoring and data collection process. These can be divided into four general
categories: those in charge of making key strategic decisions; those in charge of
particular research activities involved in monitoring; those carrying out the
monitoring activities (the “fieldwork”), and those who provide general assistance
or support across all stages. Secondly, tools for monitoring should be set in place,
linked to the specific methods chosen for each indicator. These might include
specific equipment, questionnaires, or enabling technologies. A monitoring
schedule should be established, detailing when particular measurements will be
taken. Finally, a clear data collection plan should be established, by providing
answers to the following questions:

Table 3-2. Questions to answer through the local monitoring and data collection plan

For the monitoring activities For the data collection and storage plan

What will be monitored? (includes expected Which type of data will be collected and what
outcomes and chosen indicators) is the target population or type of sample?

Where will monitoring take place? (location Who will analyse the data? (which stakeholders
of monitoring tools and data collection) or partners will perform the analyses)

Who will do the monitoring? (Stakeholders Who will store the data? (stakeholders
responsible for each type of data collection) responsible for the data platform and/or data
base)

When will monitoring take place? (Schedule How will data be presented? (how the results
– times and frequency of data collection) of monitoring will be presented to inform
policies, citizens and decision-making
processes)

88
Throughout this process, risks may arise in data collection activities, such as
delays in data collection, low response or unaffordable costs for municipalities.
Establishing risk mitigation plans before the start of data collection will make it
easier for local teams to avoid delays and inefficiencies.

STEP 5: Implementing the impact monitoring and evaluation plan

Implementing the impact evaluation, evaluating positive/negative features of


NBS impacts related to the different challenges, analysing and interpreting the
findings. Once data has been identified and collected, the next step is to analyse
and interpret it, in order to assess NBS performance in achieving established
objectives, and assess both positive and negative impacts, as well as synergies
and trade-offs. This might entail looking at results of several impact evaluation
rounds in combination as these may be relevant on the achievement of a
particular objective. If several outcomes impacts (positive and/or negative) are
considered in relation to an expected objective, the performance evaluation
should consider trade-offs and possible differences in time scales over which
indicators show that an objective has been achieved or not. Multi-criteria analysis
may be used to consider the different views of stakeholders.
The results of the data analysis should be related to the initial objectives outlined
in the theory of change. Local teams will thus be able to check whether NBS
actions have had the expected impact, or, on the contrary, have had undesired
effects. This is a good time to reflect on whether there are synergies between
outcomes, or whether there are trade-offs. As Chapter 2 underlines, in case the
results are not as expected, it is necessary to be careful when concluding that
the NBS actions are not effective. Actions may have the expected effect, but over
a longer time span.
Temporality is thus an element to consider in the global analysis of outcomes.
Some impacts (e.g., promoting social cohesion in a neighbourhood) require a
longer time to become apparent, while others can be verified almost immediately
(e.g., reducing local temperature through green walls). It is strongly
recommended to make evaluation an ongoing process, with different data
collections over time, to better assess changes.
Furthermore, conclusions should not be drawn solely based on the change in an
indicator before and after implementing the NBS, but to do a benchmarking
process where scientific standards are taken into account that indicate which
values are appropriate for an indicator (e.g., not only assess a decrease in
pollution levels after implementing an NBS, but consider when the decrease is in
line with scientific criteria). Figure 3-4 illustrates the monitoring strategy
workflow used in the EU H2020 CLEVER Cities project, to illustrate the different
stages involved in the implementation of a monitoring and data collection plan.

89
Figure 3-4. Impact Assessment process in the CLEVER cities project lifetime (Tecnalia, 2018)

STEP 6: Disseminating results and achieving policy impact

The last stage of the NBS impact assessment process involves the dissemination
of results as well as making provisions to embed them into policy practice. The
wider the dissemination, the more benefits it will have: citizens will be informed
of the activities of their local government, companies will be made aware of
business opportunities, and scientists will be able to continue advising on and
researching the best methodologies for NBS impact assessment.
We stress the importance of not only registering and reporting positive results,
tempting as that may be, but to do so for all the results obtained. Although it is
often tempting to only consider and disseminate positive effects, knowing what
has gone wrong or which parts of the implementation are susceptible to
improvement in the future are of utmost importance in order to not repeat
mistakes or waste resources by implementing the same ineffective strategies and
solutions elsewhere. It is also very important to disseminate both outcome and
process results. Reporting all results will mean that knowledge and evidence will
accumulate, benefitting everyone working with NBS.
Disseminating the knowledge generated by the local team to others not only helps
in the replicability of NBS, but also positions city councils as role model. Different
collaborative actions can be carried out to help disseminate the data, such as
scientific articles, official reports, conference presentations, talks and webinars,
or social- and mass-media interviews. It is also very helpful to create integrated
and highly visual representations of impacts, and where possible include a spatial
or GIS component to the visualization of the data, to support decision-making.
The more attractive and easier to navigate these data dissemination platforms
are, the more they will enable stakeholder collaboration and evidence-based
decision-making in the future.
The creation of NBS impact dashboards by cities or regions, which integrate GIS
technology, and allow interaction with different types of data, are gaining

90
prominence. The following image is an example of the impact dashboard created
in the city of Glasgow as part of the Connecting Nature project, as a way to map
and represent outcomes of the City’s Open Space Strategy and the impacts of
NBS implementation in different areas. The dashboard allows viewers to visualize
the interplay of different indicators (e.g. health status, social deprivation, green
space distribution) in a particular city location, and provides a flexible structure
that will be further developed as additional NBS are implemented and additional
data becomes available. It is also a useful instrument to identify types of
indicators and data that might be missing, thus orienting future impact
assessment decisions.

Figure 3-5. Glasgow City Council Dashboard (© Glasgow City Council), the Connecting Nature Project

3.3 Robust impact assessment and co-production: a necessary


relationship

The design, implementation and evaluation of nature-based solutions require the


collaboration of different stakeholders. Although the design and implementation
of monitoring and evaluation plans is often considered the part of the process
where most technical and scientific expertise is required, we argue that
monitoring and evaluation can also benefit from collaborative, co-productive
approaches. The knowledge, expertise and lived experience of many stakeholders
is relevant when deciding what outcomes to evaluate, when identifying existing
local needs, as well as when implementing monitoring strategies and gathering
relevant data. Using well-designed collaborative approaches can also reduce
costs and enhance NBS ownership, as, for example, when using citizen science
approaches to monitor biodiversity. Even for the most technical parts of
monitoring and evaluation, such as deciding on where and when to use certain
equipment for data collection, using a collaborative approach can ensure that
residents are knowledgeable of the reasons for it, and they can contribute to
equipment maintenance and/or safety. Citizen participation in monitoring and

91
evaluation efforts can enhance socially innovative solutions and accelerate the
transition to sustainability (Faivre et al., 2017).
Moreover, the multifunctional nature of nature-based solutions will mean that
different administrative departments and agencies will need to be involved in
monitoring and evaluation (Calliari et al., 2019). Monitoring NBS impacts in
different urban, rural or coastal conditions advances the knowledge acquired by
local authorities (Frantzeskaki et al., 2019). Co-production will provide
opportunities to change traditional ways of thinking and planning (Bush and
Doyon, 2019). Impact assessment might require the use of data collected and
kept in the custody of different departments, thus overcoming data and
monitoring silos. Changing traditional silo-type modes of operation, where
ecological, social and economic objectives are considered separately, the focus
needs to shift to a broader conceptualization of urban resilience and regeneration
(Dumitru et al., 2020), and to an institutional culture of cooperation (Frantzeskaki
et al., 2019). Finally, business sector stakeholders can provide valuable
information related to the economic and environmental dimensions of the NBS.
Different stakeholders help to highlight weaknesses, to prioritize interventions
and to identify the adequacy of assessment tools for diverse locations (Beceiro et
al., 2020).
The degree of stakeholder participation will depend on whether their points of
view are taken into consideration by local governments and on their proximity to
the decision-making process of interventions (Wamsler, 2017). Planners can
think of this in terms of a continuum, ranging from centralized, hierarchical
decision-making to decentralized, participatory monitoring and evaluation where
stakeholders take joint ownership of the process and are actively engaged at each
stage. Different models, or positions on this continuum, have their pros and cons.
Centralized or hierarchical decision-making models ensure a fast and potentially
less expensive process, but can be seen as poor processes by the citizens and
generate reactivity, thus undermining acceptability of different NBS strategies
and projects. On the other side of the continuum, participatory models require a
greater investment of resources (time and budget), but contribute to citizen
ownership of the solution, the creation of a culture of collaboration and
engagement, as well as a sense of community and belonging, and in the long
term might lower costs through good maintenance of the solution by the
community. Co-production approaches will also foster greater NBS-related
business opportunities through engagement with the business sector, as well as
increased network creation and trust-building.
Co-production is different from consultation or information provision, and the key
differentiating feature is that stakeholders are involved from the very beginning
in the development of monitoring and evaluation plans, in each of the steps
described in section 3.2.
We highlight five stages that are important for the co-production of impact
assessment plans. Importantly, outlining a co-production strategy and creating
specific co-production plans should happen at the very beginning of the process
of NBS design and implementation. Co-production stages are also iterative. It is
important to continuously reflect, redefine and adapt the process of monitoring
and evaluation co-production if and when needed.

92
It is also important to keep in mind that co-production is not a panacea. Ensuring
good quality co-production requires the development and strengthening of new
types of skills, resources and relationships to foster exchange and collaboration
between stakeholders. It is thus of paramount importance to take time at the
outset of the process to establish good relationships with stakeholders from the
outset, for which good communication skills and openness to multiple
perspectives is helpful. We highlight here the key stages in the planning and
implementation of an effective co-production process.

Stage1: Define the goals of, and create space for, the co-production
process
The goal of co-production of monitoring and evaluation of nature-based solutions
should be clarified from the start, by addressing questions such as: To what ends
do stakeholders need to be involved? Which amount of time needs to be allocated
to the co-production process? The goals need to be clearly communicated to
potential funders as well as participants. People are more likely to become
actively engaged when outcomes are clearly visible, and their opinions are
authentically considered and appreciated.
Answers to these questions will determine the goals that influence which actors
should be involved and in which steps of the process. Depending on the objectives
and time availability, the goals of co-production can pertain to each of the steps
outlined above, or a choice can be made to involve (different types of)
stakeholders in specific steps. For example, in the development of a theory of
change (Step 1), cities can benefit from the knowledge of the various
stakeholders to understand local needs, desires for change and how the NBS can
address them. Shared aspirations for outcomes can be formulated collaboratively
from the beginning. Other stakeholders can be involved later on in the collection
and interpretation of data (Step 5), as well as in debates and decisions on how
to adapt the NBS to improve outcomes.
Co-production requires a high amount of time and resources, openness and trust,
as well as (political) support and motivated participants. This needs to be
considered in the initial goal setting and time planning to allow and plan for
sufficient availability of time for things like initial preparation of the co-production
process, mobilisation of stakeholders or processing information for each
subsequent monitoring and evaluation step.

Stage 2: Identify and reach out to the actors that will be involved
Secondly, the actors that are sought to be involved need to be identified and
contacted. Who should be involved depends on the nature-based solution itself,
including where it is located and who is affected. It is important to explicitly go
beyond the usual suspects to guarantee greater inclusion and participation of the
weakest and give voice to critical perspectives.
Actor mapping tools facilitate the identification of suitable participants. The
Quintuple Helix approach helps identify key stakeholders across different
audiences to be targeted as part of the co-production process: 1) Academic; 2)
Industry, firms, economic system; 3) State, government, local political system;

93
4) Media-based and culture-based public – local communities, community groups,
NGO’s – mainstream and local media, environmental media; 5) Natural
environments of society – NGO’s, policy makers, political bodies, experts and
opinion leaders on NBS.

Figure 3-6. Quintuple Helix Stakeholders (adapted from Carayannis, Barth, and Campbell, 2012; Dumitru
et al., 2020)

It is important that stakeholders in each of these categories are identified early


on, and decisions are made about how they might be engaged, depending on the
objectives identified. We should not only consider the type of knowledge these
stakeholders can provide at different monitoring and evaluation stages, but also
what type of knowledge and expertise might they also acquire through this
process, how the process can contribute to building confidence among some of
the more vulnerable stakeholders, and empower them for further meaningful
participation in the implementation of the NBS.
Mobilising diverse actors requires boosting and tapping into motivation for
participation. While people might be intrinsically motivated, co-production
requires time, effort and money, and (shared) benefits might only be felt in the
long-term. Levers for motivation can include money-related complements (e.g.,
financial support, training), but also social, cultural and psychological factors
including social rewards, feeling part of a group and socialisation of the behaviour
of participation and collaboration.
Actively going out to communities and holding regular meetings that are open to
all are important conditions for enabling co-production. In addition, adequate
follow-up is essential: when participants feel that they have wasted their time,
they might become frustrated and disempowered to take up initiative in the next
stages. Each meeting and discussion stage should be followed by feedback and
the integration of issues raised into the subsequent discussions in a meaningful
way (or at least providing reasons for why particular ideas might not be possible,
or were not integrated). It is also important to monitor who does (and does not!)
benefit from the results.

94
Additionally, the different roles and responsibilities for organising the co-
production process need to be defined. Think of roles and responsibilities in terms
of process design, facilitation, aggregating the generated knowledge,
communicating results etc. The co-definition of roles and responsibilities in the
process gives clarity about what is expected from actors and helps them feel
comfortable in and adopting their (new) roles and functions.
One of the challenges of co-production is balancing all the interests and needs.
For example, each stakeholder might have a different vision about the objectives
to be set in the city’s theory of change. Inclusive co-production means that the
process format is based on mutuality, reciprocity and equality between different
groups (e.g., experts, citizens), for example in terms of considering capabilities
and time restrictions of different groups and giving equal voice to everyone.
Communication and engagement need to consider the different capabilities,
values, languages and resources of participants, as well as potential pre-existing
cooperation or contestation between actors and institutional power structures.
Ideally, this allows for open discussion and sharing of opinions in a joint learning
setting, which builds on the recognition that different views are not exclusive but
complementary.

Stage 3: Plan the co-production activities and tools


Thirdly, the co-production activities have to be planned with a timeline of when
these are going to happen. The main question to be addressed here is ‘how’,
relating to the right type of formats and tools to engage with the stakeholders.
For example: How should different actors be involved in the construction of a
theory of change? How will they be involved in the selection of indicators and
data collection?
Specific co-production tools facilitate each step of the process towards desired
goals. Tools are highly diverse. The choice of tools depends on the goals of the
co-production process, on the specific impact monitoring and evaluation step,
and on the type of actors involved. For example, visioning exercises serve to
generate inspiring future images and ideas; they are particularly useful at the
beginning to support the development of a theory of change, as well as to align
diverse actors and to create long-term, systemic and normative aspirations.
Citizen science approaches can support wide data generation, but need to be
complemented with workshops for joint reflection upon the data.
Citizen science refers to public participation in scientific research and projects,
not only to collaborate with scientists collecting data but also has the potential to
engage the public in research by modifying the knowledge, attitudes and
behaviour of citizens (Peter et al., 2019). This participatory research can promote
the efficiency and effectiveness of research processes, as well as foster social
inclusion, empowerment and sustainability (van de Gevel et al., 2020). Citizen
participation through citizen science can provide a wealth of data to create
evidence that can address real-world problems, which would otherwise be
insurmountable for small teams of professionals (Gildefer et al., 2019)
Performing a classification of citizen science projects, linked to voluntary forms
of participation, Follet and Strezov (2015), grouped these projects into: a)
contributory projects: citizens participate in data collection and analysis, as well

95
as in the dissemination of results; b) collaborative projects: in addition to the
previous functions, the participants would help in the design of the study and
interpretation of the data and conclusions; c) co-created projects: collaboration
would be carried out at all stages of the project, from the development of
hypotheses to the discussion of results, and the answer to new research
questions. Therefore, in the monitoring of the NBS, citizens can be involved from
the co-design of the strategic objectives of the local authorities, until the last
phases of data collection and transfer of results.
Although citizen science approaches have a lot of potential, they are not
appropriate for all types of outcomes assessed, especially those for which specific
expertise is required (Wamsler et al., 2020). Although the data collected by
citizens may sometimes have levels of accuracy similar to the data collected by
experts, participants need to be engaged for long time periods in larger groups
and with specific training (Aceves-Bueno et al., 2017).
The co-production activities and tools need to be planned from the outset,
following along the steps for impact assessment and monitoring, but also
considering that the process will likely need to change and adapt.
After selecting the co-production tools, it is important to identify the materials,
skills and other requirements needed to implement the tool. Think for example
of the space/room, atmosphere and time needed.

Stage 4: Reflect on the co-production process and results


Co-production processes are never set in stone. They are open processes and
evolve over time as learning progresses. They ‘go with the flow’ of the
participants’ ideas and needs. This requires continuous reflexivity. Reflexivity
helps to identify lessons learned and to adapt the process in light of (changing)
objectives. Which goals does the process aim to achieve? Is the process on the
way to achieve these, or do we need adaptations? Reflexive monitoring can help
to achieve reflexivity (see section 3.4.1).

Stage 5: Communicate about the co-production process


The co-production process and results need to be politically and societally known
and accepted. This closely links to Step 6 (dissemination of results and achieving
policy impact) of the impact monitoring and evaluation plans. This can be
achieved through outreach and awareness raising activities such as campaigns
and public events. Communication formats should be accessible, tailored to and
inclusive of different target audiences, use innovative techniques (e.g.,
storytelling, puppet play, etc.), tell an inspiring story and clearly articulate the
results. The participants of the process can be actively engaged in such activities.
If the evaluation and monitoring process is broadly known, greater collaboration
can be achieved and thus obtain data from more sources, therefore, co-operation
with the media can help disseminate the importance of evaluation. Finally,
science-practices partners (i.e., universities, research institutes, etc.) serve as
guides in cities to carry out each of the steps of the process. Academic entities
can establish synergistic collaborations with cities, being able to use the
evaluation results to disseminate them internationally, and to accumulate more

96
evidence on the NBS. Successful approaches can then be transferred between
case studies, communities and countries (Raymond et al., 2017a), with the
support of the established networks.

3.4 Innovative tools for monitoring and evaluation of nature-based


solutions

Monitoring and evaluation of nature-based solutions can benefit significantly from


technology-supported innovations. Collaborative technological approaches have
been encouraged (Ershad-Sarabi et al., 2019), and the existence of new
platforms that facilitate co-production and interaction between citizens and
governments, especially in the context of urban development, has been
highlighted (Falco and Kleinhans, 2018). We provide a few examples of innovative
methodologies for monitoring and evaluation: a collaborative approach to
enhance structured reflection and reflexivity regarding monitoring and
evaluation; an online tool to create robust monitoring and evaluation plans; and
a smartphone-supported, automatized data collection and citizen engagement
tool.

3.4.1 Reflexive monitoring - Connecting Nature project

Reflexive monitoring is a participatory and dynamic monitoring and learning


process that enables practitioners to gain insight into the progress and direction
of their nature-based solution project in real time, and not only retrospectively.
Reflexive monitoring stimulates learning, supports the identification of barriers
and opportunities and enables flexible responses to changing circumstances and
objectives. It is about adopting a reflexive mind set: reflexivity is the ability to
interact with and alter the environment within which one operates. This allows
practitioners to take actions that influence the context in which they work for the
implementation of their nature-based solution. It is a particularly useful process
for the nature-based solution core project team, although it can be adapted to
involve and stimulate reflexivity among a wider range of stakeholders.
Reflexive monitoring can help for example with continuous reflection about
whether indicators fit the outcomes and goals of the project or whether they need
adaptation, or the appropriateness of data and data collection. It can also support
reflection about the process itself, including whether there needs to be more time
for co-production, or whether the right stakeholders are involved.
Within the H2020 Connecting Nature project, the innovative reflexive monitoring
tool has supported cities in reflecting on their progress in the planning, delivery,
evaluation, and stewardship of NBS, being able to record what actions allowed
them to overcome the difficulties encountered. The following section is based on
the Reflexive Monitoring Guidebook by Lodder et al (2020). The Connecting Nature
cities of Genk (Belgium), Glasgow (Scotland) and Poznan (Poland) have found it
is wise to reserve space and time to become familiar with the steps and the tools
before proceeding with them. Once the reflexive monitoring process is aligned
with your daily activities, you will be able to identify the benefits and act on what
you learn.

97
For the Connecting Nature cities, a six-step procedure (see Figure 3.7) has been
developed to implement the reflexive monitoring process. These steps can be
applied in parallel to the steps for developing impact monitoring and evaluation
plans. Reflexive monitoring should accompany all the steps outlined for robust
impact assessment.

Figure 3-7. Steps in the reflexive monitoring process with accompanying tools (source: Lodder et al., 2020)

The reflexive monitoring process outlined below is supported by seven reflexive


monitoring tools which may be applied by NBS practitioners. The tools are based
on a selection of the tools presented in the Reflexive Monitoring in Action
guidebook by Van Mierlo et al (2010).

RM step 1: Rethink what learning process you need to achieve the goals
of the Nature-based Solution
When describing the process of co-production, we stressed the importance of
clearly defined co-production goals. Beyond the goals of the nature-based
solution, and the process of co-production, we also recommend identifying clear
learning goals for the different actors involved. It includes how the process of
NBS design and implementation is different from other planning processes, and
the different departments that need to be involved. Next, it is important to
acknowledge that reflexive monitoring is a novel process for all actors involved.
For it to be successful you need to plan for space and time to get acquainted with
the tools and to include them into your daily activities.

98
RM step 2: Define the roles within the project team
From the very outset of the reflexive monitoring process, it should be made clear
that each actor has a role in the process and that exercising this role will involve
collaborating closely and meeting regularly. The level of involvement of each one
depends on the steps in the process.

RM step 3: Start with recording important events and translate them into
your dynamic learning agenda
Start with recording a timeline of events during one or two months. This is to
trace important moments, insights, events, that influence the development of the
impact monitoring and evaluation plan. Discuss the timeline of events with your
project team and distil important moments in time where something changed
that helped or hindered to process. Include the critical turning points to your
dynamic learning agenda and add learning questions and follow-up questions for
each turning point. This allows for collective reflection on the essence and
difficulty of the challenges that are dynamic and change over time. The objective
of the dynamic learning agenda is to link long-term aims and learning objectives
to concrete actions in the short term. By formulating, recording and tracking
challenges in time the learning journey itself can be evaluated as a dynamic
process.

RM step 4: Use learning sessions to identify learning outcomes


This step is about supporting the team to improve the learning process and
analyse the outcomes. To facilitate this, we recommend the organising of learning
sessions with the reflexive monitoring team. During the learning sessions each
newly added item on the dynamic learning agenda is discussed. The critical
turning points in the development of the project and learning questions are
discussed and if needed reformulated to increase their reflexivity. After all items
on the dynamic learning agenda are discussed, the expert and team identify
learning outcomes. Learning outcomes are innovative ways the team handles the
barriers or opportunities captured in the dynamic learning agenda.
We operationalized a framework for reflexive learning outcomes based on Beers
and Van Mierlo (2017) that distinguish between the following categories: (1)
Rules guiding actors’ practices, for example tendering procedures or the way a
city department is organised; (2) Relations between actors and between the
nature-based solution and its context, for example who is involved in the planning
process; (3) Practices concerning common ways of working, for example how the
team collaborates internally; and (4) Discourse related to the future of the
nature-based solutions, for example the way a mayor talks about the benefits of
nature-based solutions for the city. Analysing learning outcomes in detail helps
the team to better understand and explain to others what they learnt, identify
remaining gaps in knowledge that can be covered through additional stakeholder
collaboration or training and capacity building exercises, and highlight
innovations in urban planning, including the monitoring and evaluation dimension
of NBS.

99
RM step 5: Communicate about the reflexive monitoring process to peers
and project outsiders
Reflexive monitoring is a novel governance process that allows many lessons to
be learned. It is valuable to share these lessons, along with tips and tricks, with
other actors who might benefit from the method. The following two tools are
selected to support this exchange: the eye-opener workshop and the personal
learning narrative. The purpose of eye-opener workshops is to share what is
learned from co-producing nature-based solutions with people who are not yet
involved in your project. For example, colleagues from other departments, the
mayor’s office or professionals working with co-production or involved in nature-
based solutions projects. Personal learning narratives are stories that describe
the learning journey of yourself or your team members throughout the co-
production process. These may take the form of an experience, a hindering factor,
a struggle or a challenge. These personal stories can be shared in different ways
to supplement regular reports. For example, a participant records a video about
his or her own learning journey and it is shared through social media or played
at an eye-opener workshop.

RM step 6: Reflect upon reflexive monitoring as a method for knowledge


generation regarding how to educate about the multiple
benefits of nature-based solutions and how to adapt the
planning process in real-time
In step six, sessions can be organised to reflect upon the effectiveness of the
reflexive monitoring method itself and compare and share the learning outcomes.
These sessions give practitioners the chance to share their experience of working
through the various steps and using the tools of the method, which may in turn
be adapted based on the feedback received or changing needs. Peer-to-peer
learning events can be used for the sharing and comparing of the learning
outcomes of different teams. Think of organising sessions to learn how others
dealt with similar challenges and barriers, sharing personal learning narratives
and celebrating innovations to inspire each other.

3.4.2 iAPT (Impact Assessment Planning Tool) - Connecting Nature project


Developed within the Connecting Nature Project, iAPT is intended to be a
decision-support tool for cities to create their NBS evaluation and monitoring
plans. The main objective is that users, mostly urban planners, can obtain their
individualized monitoring and evaluation plan adapted to the characteristics of
their location, online, easily and intuitively.
The tool supports planners and project teams to go through an abbreviated
version of the step by step process described at the beginning of this chapter.
After users indicate some characteristics of the location, placing it on an
interactive map, they outline their theory of change, by reflecting on the
characteristics of their NBS and explicitly relating them to certain outcomes, by
choose from a list of possible impacts grouped into different impact categories
(e.g., health and wellbeing, social cohesion, greenspace management, etc.).

100
Once users have made their initial selection of benefits, iAPT provides suggestions
regarding relevant indicators to assess identified expected outcomes. Users will
be able to consult a series of factsheets regarding methodologies for particular
indicators to get a better idea of what they represent and what methods and
measurements can be used for them. While users will select which indicators to
measure, iAPT will suggest other indicators that are equally important and might
not have been considered by the project team, to create a coherent impact
assessment framework that reflects the multifunctional character of nature-based
solutions.
Subsequently, iAPT will offer various methodological options for each of the
indicators. As explained in this chapter, users must make the choice considering
three criteria: data quality, temporal adequacy, and the cost-benefit ratio. The
tool will be connected to the recently launched Connecting Nature-Based
Enterprise platform, to suggest nature-based enterprises or experts that provide
support or services for a given monitoring and evaluation step or component.
Finally, users will be able to obtain and download a specific assessment plan for
their NBS, adapted to their location. This plan will contain the selected indicators,
how to measure them, as well as supplementary material and methodological
recommendations. Users can carry out the customization process as many times
as they deem convenient. Future developments of this tool could link the
evaluation plans with real data of the indicators, to complete the whole process
of data analysis and help in the dissemination of results.

3.4.3 Urban GreenUP Tool - Urban GreenUP project


As part of the monitoring strategy of the city of Valladolid, a smartphone
application has been developed by GMV, within the Urban GreenUP Project. This
is an example of an innovative technology-supported data collection platform,
conceived to act as another sensor for the monitoring program of the city, and
track both the interest generated by the NBS in citizens, as well as to assess the
use of the Green Corridor. The application will allow the collection of various
interrelated data relating to a specific user (with an identified profile). Some of
these data are collected automatically, by leveraging Smartphone sensor
(positioning by GPS/BT; position and time spent in an NBS), and others will be
actively filled in by the user (surveys, ratings). All the information provided by
the users is treated anonymously.
The smartphone application is also designed to raise awareness and increase
nature-based solutions engagement, showing a notification if it detects that the
user is near a relevant location, and providing information regarding the purpose
of the deployed or planned NBS. It can contribute to data collection for the
following challenges: Green space management (Sustainability of green areas;
Quality of life for elderly people; Perceptions of connectivity and mobility;
Recreational cultural value); Participatory planning and governance (Perceptions
of citizens on urban nature); Social justice and social cohesion (Green intelligence
awareness); Public health and well-being (Increase in walking and cycling in and
around areas of interventions).

101
Figure 3-8. URBAN GreenUP tool (Source: GMV-S).
Acknowledgements: Fátima López Mateos, Jesús Ortuño Castillo [GMV, URBAN GreenUP partners], Alicia
Villazán Cabrero [Valladolid City Council, URBAN GreenUP partners and front-runner city]

Moreover, the smartphone application promotes the use of the green corridor
throughout scoreboards and gamification. A scoreboard can serve to motivate the
users through the use of rankings, or by providing information on usage scores
in general. It also serves as a vehicle for promotions and discounts related to the
NBS. The information will be sent to a server platform that will store the actions
and information provided by the users (location and information). Data collected
will be used to calculate some of the indicators for the Valladolid monitoring
program. Currently, the use of the App and data beyond the European project is
not foreseen, but could be an option to consider in the future. For the
municipality, this data collection is important not only in terms of assessing the
impact of the URBAN GreenUP project as a whole but also as an indicator of the
degree of citizen acceptance of the re-naturalization actions implemented by the
City Council.
The application will allow the collection of various interrelated data relating to a
specific user (with an identified profile). Some of these data are collected
automatically (position and time spent in an NBS), and others will be actively
filled in by the user (surveys, ratings). The information provided by citizens when
completing their profile is used to segment the results providing data for
monitoring and evaluation by social groups. This segmented analysis of how each
social profile uses and perceives NBS can be applied in the design of future urban
re-naturalization plans
This monitoring system is a considerable improvement over more traditional
monitoring methods. As a main advantage, the use of these technologies
encourages the interaction of citizens and their participation in the design of their
own town. As a drawback, it should be noted that the population sample studied
is only that which handles these technologies and maybe a non-representative
population sample.

102
Although the app is not open source and has been specifically designed for
Valladolid city and their specific NBS actions, functionalities can be adapted to
other cities.

3.5 Conclusions
Throughout this chapter, the importance of developing robust evaluation and
monitoring plans has been emphasized, to assess the processes, outputs and
outcomes involved in NBS design and implementation. Also highlighted in this
chapter is the idea that NBS impact assessment should not be conducted in
isolation by local authorities, but must have the support and active collaboration
of multiple stakeholders such as scientists, companies, media, citizens and policy
makers. The closer local teams are to the co-production end of the continuum,
the richer, more effective and less costly impact assessment will be, while
acceptability, empowerment of vulnerable groups and the creation of a culture of
NBS evaluation will also be fostered.
Monitoring and evaluation in cities and regions can also have a clear educational
role, since it is possible to learn from mistakes and disseminate successes
(Pappalardo and La Rosa, 2020). Evaluation contributes to the development of
long-term plans and goals for NBS (Kabisch et al., 2016), and leads to new
insights and active learning, including failures, to improve future implementations
(Connop et al., 2016). Impact assessment should be carried out across multiple
categories of impacts, and synergies between outcomes should be considered, as
well as NBS evolution over time (Calliari et al., 2019).
Throughout this handbook, you will find descriptions of many different European
NBS projects and their monitoring and evaluation frameworks and strategies.
They illustrate the step-by-step approach outlined at the beginning of this
chapter, and are examples of different co-production strategies for monitoring
and evaluation. Many of the difficulties encountered revolved around the lack of
an evaluation culture on at local levels, which resulted in monitoring and
evaluation not being planned from the beginning, as well as to many
misconceptions about indicators, methodologies, costs and efforts. Collaboration
between scientists, technical experts, municipalities and other stakeholders
contributed to overcoming these barriers and advancing knowledge on conditions
for successful and robust impact evaluation for nature-based solutions. Lessons
from all these projects have been captured in the principles and approaches
described here.
The ultimate goal of the process of creating robust impact assessment plans on
a local level is to gather long-term robust evidence regarding NBS performance
in particular spatial contexts and for different social groups, and to embed this
evidence to support smart policy decisions to foster sustainability, wellbeing, and
resilience (Dumitru et al., 2021). By establishing a culture of periodic evaluation,
local authorities will be able to learn with each intervention and get as close as
possible to achieving their strategic goals and building sustainable and socially
just environments.

103
3.5 References

Aceves-Bueno, E., Adeleye, A.S., Feraud, M., Huang, Y., Tao, M., Yang, Y., and Anderson, S. E., ‘The accuracy
of citizen science data: a quantitative review’, Bulletin of the Ecological Society of America, Vol. 98,
No 4, 2017, pp. 278-290.

Beceiro, P., Brito, R.S., and Galvão, A., ‘The Contribution of NBS to Urban Resilience in Stormwater
Management and Control: A Framework with Stakeholder Validation’, Sustainability, Vol. 12, No 6,
2020, Art. no 2537.

Beers, P.J. and van Mierlo, B., ‘Reflexivity and learning in system innovation processes’, Sociologia Ruralis,
Vol.57, No 3, 2017, pp. 415-436.

Brandsen, T. and Honingh, M., ‘Distinguishing different types of coproduction: A conceptual analysis based on
the classical definitions’, Public Administration Review, Vol. 76, No 3, 2016, pp. 427-435.

Bush, J. and Doyon, A., ‘Building urban resilience with nature-based solutions: How can urban planning
contribute?, Cities, Vol. 95, 2019, Art. no 102483.

Calliari, E., Staccione, A., and Mysiak, J., ‘An assessment framework for climate-proof nature-based solutions’,
Science of the Total Environment, Vol. 656, 2019, pp. 691-700.

Carayannis, E.G., Barth, T.D., and Campbell, D.F., ‘The Quintuple Helix innovation model: global warming as
a challenge and driver for innovation’, Journal of Innovation and Entrepreneurship, Vol. 1, No 1, 2012,
pp. 1-12.

Cardno, Toolkit for Monitoting and Evaluation Data Collection, Pacific Women Shaping Pacific Development,
Cardno Emerging Markets, Brisbane, 2017.

CARE Emergency Group, Monitoring and Evaluation Toolkit, CARE Emergency Group, 2017. Retrieved from
https://www.careemergencytoolkit.org/management/9-monitoring-and-evaluation/

CLES, Evaluating regeneration projects and programmes, Centre for Local Economic Strategies, Manchester,
2010.

Cohen-Shacham, E., Andrade, A., Dalton, J., Dudley, N., Jones, M., Kumar, C., Maginnis, S., Maynard, S.,
Nelson, C.R., Renauld, F.G., Welling, R., and Walters, G., ‘Core principles for successfully
implementing and upscaling Nature-based Solutions’, Environmental Science and Policy, Vol. 98,
2019, pp. 20-29.

Compass, How to Develop a Monitoring and Evaluation Plan, Springboard Compass, 2010. Retrieved from:
https://www.thecompassforsbc.org/how-to-guides/how-develop-monitoring-and-evaluation-plan

Connop, S., Vandergert, P., Eisenberg, B., Collier, M.J., Nash, C., Clough, J., and Newport, D., ‘Renaturing
cities using a regionally-focused biodiversity-led multifunctional benefits approach to urban green
infrastructure’, Environmental Science and Policy, Vol. 62, 2016, pp. 99-111.

Creswell, J.W. and Creswell, J.D., Research design: Qualitative, Quantitative, and Mixed Methods Approaches,
SAGE Publications, Thousand Oaks, 2018.

Djenontin, I.N.S. and Meadow, A.M., ‘The art of co-production of knowledge in environmental sciences and
management: lessons from international practice’, Environmental Management, Vol. 61, No 6, 2018,
pp. 885-903.

Dumitru, A., Frantzeskaki, N., and Collier, M., ‘Identifying principles for the design of robust impact evaluation
frameworks for nature-based solutions in cities’, Environmental Science and Policy, Vol. 112, 2020,
pp. 107-116.

Dumitru, A., Tomé-Lourido, D., Young, C., Connop, S., Rhodes, M.L., Dick, G., Sermpezi, R., Impact
Assessment Guidebook: Developing robust monitoring and evaluation plans for nature-based
solutions, Connecting Nature Grant Agreement number 730222, 2021.

Ershad Sarabi, S., Han, Q., Romme, A.G.L., de Vries, B., and Wendling, L., ‘Key enablers of and barriers to
the uptake and implementation of nature-based solutions in urban settings: a review’, Resources, Vol.
8, No 3, 2019, Art. no 121.

European Commission, Indicative Guidelines on Evaluation Methods: Monitoring and Evaluation Indicators,
Directorate-General: The New Programming Period 2007-2013, 2006.

Faivre, N., Fritz, M., Freitas, T., de Boissezon, B., and Vandewoestijne, S., ‘Nature-Based Solutions in the EU:
Innovating with nature to address social, economic and environmental challenges’, Environmental
Research, Vol. 159, 2017, pp. 509-518.

Falco, E. and Kleinhans, R., ‘Digital participatory platforms for co-production in urban development: A
systematic review’, International Journal of E-Planning Research, Vol. 7, No 3, 2019, pp. 52-79.

104
Follett, R. and Strezov, V., ‘An analysis of citizen science based research: usage and publication patterns’,
PloS One, Vol. 10, No 11, 2015, Art. no e0143687.

Frantzeskaki, N. and Kabisch, N., ‘Designing a knowledge co-production operating space for urban
environmental governance—Lessons from Rotterdam, Netherlands and Berlin, Germany’,
Environmental Science and Policy, Vol. 62, 2016, pp. 90-98.

Frantzeskaki, N., McPhearson, T., Collier, M.J., Kendal, D., Bulkeley, H., Dumitru, A., Walsh, C., Noble, K.,
van Wyk, E., Ordóñez, C., Oke, C., and Pintér, L., ‘Nature-based solutions for urban climate change
adaptation: linking science, policy, and practice communities for evidence-based decision-making’,
BioScience, Vol. 69, No 6, 2019, pp. 455-466.

Galuszka, J., ‘What makes urban governance co-productive? Contradictions in the current debate on co-
production’, Planning Theory, Vol. 18, No 1, 2019, pp. 143-160.

Geneletti, D., Zardo, L., and Cortinovis, C., ‘Promoting nature-based solutions for climate adaptation in cities
through impact assessment’, Handbook on biodiversity and ecosystem services in impact assessment,
Edward Elgar Publishing, Cheltenham, 2016, pp. 428-452.

Gilfedder, M., Robinson, C.J., Watson, J.E., Campbell, T.G., Sullivan, B.L., and Possingham, H.P., ‘Brokering
trust in citizen science’, Society and Natural Resources, Vol. 32, No 3, 2019, pp. 292-302.

Kabisch, N., Frantzeskaki, N., Pauleit, S., Naumann, S., Davis, M., Artmann, M., Haase, D., Knapp, S., Korn,
H., Stadler, J., Zaunberger, K., and Bonn, A., ‘Nature-based solutions to climate change mitigation
and adaptation in urban areas: perspectives on indicators, knowledge gaps, barriers, and opportunities
for action’, Ecology and Society, Vol. 21, No 2, 2016, Art. no 39.

Kates, R.W., Clark, W.C., Corell, R., Hall, J.M., Jaeger, C.C., Lowe, I., McCarthy, J.J., Schellnhuber, H.J., Bolin,
B., Dickson, N.M., Faucheux, S., Gallopin, G.C., Grübler, A., Huntley, B., Jäger, J., Jodha, N.S.,
Kasperson, R.E., Mabogunje, A., Matson, P., Mooney, H., Ill, B.M., O’Riordan, T., and Svedin, U.,
‘Sustainability Science’, Science, Vol. 292, No 5517, 2001, pp. 641-642.

Lodder, M., Allaert, K., Hölscher, K., Notermans, I., and Frantzeskaki, N., Reflexive Monitoring Guidebook:
using continuous evaluation techniques to adapt your nature-based solution planning process in real-
time, Connecting Nature Grant Agreement number 730222, 2020.

Nesti, G., ‘Co-production for innovation: the urban living lab experience’, Policy and Society, Vol. 37, No 3,
2018, pp. 310-325.

Pappalardo, V. and La Rosa, D., ‘Policies for sustainable drainage systems in urban contexts within
performance-based planning approaches’, Sustainable Cities and Society, Vol. 52, 2020, Art. no
101830.

Peter, M., Diekötter, T., and Kremer, K., ‘Participant outcomes of biodiversity citizen science projects: a
systematic literature review’, Sustainability, Vol. 11, No 10, 2019, Art. no 2780.

Raymond, C.M., Berry, P., Breil, M., Nita, M.R., Kabisch, N., de Bel, M., Enzi, V., Frantzeskaki, N., Geneletti,
D., Cardinaletti, M., Lovinger, L., Basnow, C., Monteiro, A., Robrecht, H., Sgrigna, G., Munari, L., and
Calfapietra, C., An impact evaluation framework to support planning and evaluation of nature-based
solutions projects. Report prepared by the EKLIPSE Expert Working Group on Nature-Based Solutions
to Promote Climate Resilience in Urban Areas, Centre for Ecology and Hydrology, Wallingford, 2017a.

Raymond, C.M., Berry, P., Breil, M., Nita, M.R., Kabisch, N., de Bel, M., Enzi, V., Frantzeskaki, N., Geneletti,
D., Cardinaletti, M., Lovinger, L., Basnow, C., Monteiro, A., Robrecht, H., Sgrigna, G., Munari, L., and
Calfapietra, C., ‘A framework for assessing and implementing the co-benefits of nature-based solutions
in urban areas’, Environmental Science and Policy, Vol. 77, 2017b, pp. 15-24.

United Nations, Handbook on Planning, Monitoring and Evaluating for Development Results, United Nations,
New York, 2010.

United Nations, Transforming our World: The 2030 Agenda for Sustainable Development. United Nations, New
York, 2015.

van De Gevel, J., van Etten, J., and Deterding, S., ‘Citizen science breathes new life into participatory
agricultural research. A review’, Agronomy for Sustainable Development, Vol. 40, No 5, 2020, pp. 1-
17.

van Mierlo, B.C., Regeer, B., van Amstel, M., Arkesteijn, M.C.M., Beekman, V., Bunders, J.F.G., de Cock
Buning, T., Elzen, B., Hoes, A.C., and Leeuwis, C., Reflexive Monitoring in action. A guide for
monitoring system innovation projects, Communication and Innovation Studies, WUR; Athena
Institute, VU, Wageningen/Amsterdam, 2010.

Wamsler, C., ‘Stakeholder involvement in strategic adaptation planning: Transdisciplinarity and co-production
at stake?’, Environmental Science and Policy, Vol. 75, 2017, pp. 148-157.

105
Wamsler, C., Alkan-Olsson, J., Björn, H., Falck, H., Hanson, H., Oskarsson, T., Simonsson, E., and Zelmerlow,
F., ‘Beyond participation: when citizen engagement leads to undesirable outcomes for nature-based
solutions and climate change adaptation’, Climatic Change, Vol. 158, No 2, 2020, pp. 235-254.

Xing, Y., Jones, P., and Donnison, I., ‘Characterisation of nature-based solutions for the built environment’,
Sustainability, Vol. 9, No 1, 2017, Art. no 149.

Zorita, S., García-Pérez, I., Murphy-Evans, N., Rödl, A., and Barone, E., Deliverable 4.3 Monitoring strategy
in the FR interventions, Clever Cities, Grant Agreement number 776604, 2020.

106
Drawing on knowlegde from projects
funded by the European Union

CLEVER Cities
Hamburg (DE) London (GB) Milan (IT)
Belgrade (RS) Larissa (GR) Madrid (ES) Malmö (SE) Sfântu Gheorghe (RO)

CLEVER Cities aims to drive a new kind of nature-based urban transformation for sustainable
and socially inclusive cities across Europe, South America and China. Its local teams including
citizens, businesses, knowledge partners and local authorities are co-creating nature-based
interventions in Hamburg, London and Milan to regenerate cities, improve the environment,
generate economic opportunities and make deprived urban districts healthier places to live.
Through multi-disciplinary learning, exchange and collaboration with Fellow cities Belgrade,
Larissa, Madrid, Malmö, Sfântu Gheorghe and Quito, the project is developing a CLEVER Solu-
tions Basket with innovative technological, business, financing and governance solutions to
adapt nature-based interventions for the needs of towns and cities around the world.

Image: Urban Innovation Partnership Meeting Hamburg - Photo © Asja Caspari


SCOPE Fostering sustainable, socially inclusive urban regeneration through nature

Approach to Impact Assessment Main Challenges addressed

The decision-making process for the develop-


ment of the project’s monitoring framework
1. Climate Resilience
was iterative and collaboratively designed with
Front-runner cities and stakeholders involved 2. Water Management
in their local Urban Innovation Partnerships
(UIPs). A first framework to guide local impact 3. Natural and Climate Hazards
assessment processes was developed using a
Theory of Change model. The second phase in- 4. Green Space Management
volved cross-comparing the Theory of Change
model against the baseline data of each city, 5. Biodiversity
then conducting a SMART model analysis in or-
der to prioritize the most salient themes for im- 6. Air Quality
pact monitoring. Afterwards, Local Monitoring
Plans were developed for each city based on 7. Place Regeneration
four macro-areas of indicators, namely: envi-
ronmental, human health and well-being, sa- 8. Knowledge and Social Capacity Building
fety and security, and economic prosperity. For
each thematic area, a performance model was 9. Participatory Planning and Governance
developed for identifying who is doing what,
how, with which tools and at what point of the 10. Social Justice and Social Cohesion
project’s lifetime.
11. Health and Wellbeing

12. New Economic Opportunities & Green Jobs


Involved Stakeholders and roles

All relevant stakeholders are integrated in the Lessons learned


process of co-defining the monitoring KPIs, in-
cluding strategic leads, operational leads, tech- In order to apply Theory of Change models to
nical and academic advisors and community monitoring processes, technical support is nee-
members. A highly collaborative approach was ded to help cities identify the outcomes and im-
developed between thematic experts in the pro- pacts that they expect from NBS. The project
ject and local monitoring teams to coordinate team found it challenging to define monitoring
the KPIs co-development and data gathering. KPIs, especially those related to social out-
By emphasizing the importance of community comes such as health and wellbeing or social
building, the project has created the necessary cohesion. Iterative feedback from thematic ex-
conditions for potential co-management of NBS perts was required to help cities overcome this
by citizens. challenge. This highlights the need of including
a robust scientific methodology in the process
of co-defining KPIs. For urban regeneration
projects that expect to monitor NBS co-bene-
Municipal Administrations fits to well-being and health, it is key to create
community-driven processes and consider sta-
Regional/national statistics authority keholders’ different expectations.

Citizen

Scientists / Academia

NGOs

Schools and Kindergartens Learn more


www.clevercities.eu
Housing Associations

The CLEVER Cities project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 776604
Drawing on knowlegde from projects
funded by the European Union

proGIreg
productive Green Infrastructure for post-industrial
urban regeneration with and for citizens
Dortmund (DE) Turin (IT) Zagreb (HR) Ningbo (CN)
Cascais (PT) Cluj-Napoca (RO) Piraeus (GR) Zenica (BA)

ProGIreg uses nature for urban regeneration with and for citizens. The project is funded by the
European Commission under the Horizon 2020 programme and runs from June 2018 until 2023.
In proGIreg’s front-runner cities’ Living Labs, eight different nature-based solutions (NBS) are
harnessed to create productive green infrastructure that not only helps improve living conditions
and reduce vulnerability to climate change, but also provides measurable economic benefits to
citizens and entrepreneurs in post-industrial urban districts. The follower cities learn from the
front runners through mutual exchange and replicate successful approaches. All the work done
in the Living Labs is characterized by an inclusive approach, whereby local citizens, governments,
businesses, NGOs, and universities co-create the nature-based solutions together, from planning
to implementation. To ensure replication beyond the project cities, proGIreg develops self-sustai-
ning business models for nature-based solutions, based on scientific assessment of the multiple
benefits they provide for social, health, ecological, and economic regeneration.

Image: Community work at Mirafiori urban gardens in Turin - Photo © Federica Borgato and Umberto Costa
SCOPE making NBS productive for regeneration at district level

Approach to Impact Assessment Main Challenges addressed

The impact of the implemented NBS is evaluated over


four assessment domains: social aspects, health, en-
1. Climate Resilience
vironment and economy. Benefits are evaluated at
both district and NBS level. At the district level, spa- 2. Water Management
tial data from existing administrative databases and
GIS-derived data are used to evaluate indicators in 3. Natural and Climate Hazards
the four domains all along the project, on a yearly
basis. A general population survey aimed at collec- 4. Green Space Management
ting data on social, health, and economic indicators
at the district level is performed before and after the 5. Biodiversity
implementation of the NBS and compared with ana-
logous results obtained in a control district, having 6. Air Quality
similar characteristics with respect to the Living Lab,
but where no NBS (or minimal NBS) are planned. Ten 7. Place Regeneration
tools and specific monitoring plans have been deve-
loped to monitor the impact of the single NBS (e.g., 8. Knowledge and Social Capacity Building
life-cycle assessments, NBS-users’ questionnaires,
or observational tools), taking into account cost-ef- 9. Participatory Planning and Governance
fectiveness and gathering comparable data.
10. Social Justice and Social Cohesion

11. Health and Wellbeing


Involved Stakeholders and roles 12. New Economic Opportunities & Green Jobs

In proGIreg, the so-called quadruple-helix model has


been adopted throughout the project, from co-de-
sign to impact evaluation. The quadruple-helix ap- Lessons learned
proach represents the core team in each Living Lab
consisting of four key stakeholder groups: civil so- The planning of monitoring activities should closely
ciety (NGOs and individual citizens), academia (uni- involve researchers, local administration and peo-
versities and research institutions), governmental ple responsible for the data collection, since plan-
institutions (local governments and other public aut- ned activities should take into account administra-
horities) and the private sector (especially SMEs). In tive barriers and availability of sufficiently trained
the development of the impact evaluation, the ap- staff. Moreover, co-designing the monitoring acti-
proach has resulted in collaboration with a broad va- vity could help focusing on the real expectations of
riety of actors: administrative support, coordination the involved population. To avoid potential pitfalls,
and data collection from local authorities; scientific stakeholders should get involved at an early stage,
support in planning and conducting the monitoring which also ensures scientific robustness and social
activities, and interpretation of the outcomes with significance of the collected data. The monitoring
academic partners; advice on the economic and in- activities should be cost-effective in correlation
novation impact and support in data collection from with expected results. In case the implemented
the industry; social engagement and support in data NBS are not suitable to produce effects that can
collection through citizen science approaches. be evaluated by the recommended indicators (for
instance, because the NBS is too small or close
to several other NBS), specific NBS-level tools and
Municipal Administrations appropriately scaled indicators need to be deve-
loped. This will ensure data reliability, but limits
Scientists / Academia cross-site comparability.

NGOs

Green Businesses
Learn more
www.progireg.eu
Schools and Kindergartens

The proGIreg project has received funding from the European Union’s Horizon 2020
innovation action programme under grant agreement no. 776528.
This work was financially supported by the National Key Research and
Development Programme of China (2017YFE0119000).
EdiCitNet
Edible Cities Network: Integrating Edible City Solutions
for social, resilient and sustainably productive cities
Andernach (DE) Berlin (DE) Havana (CU) Oslo (NO) Rotterdam (NL)
Carthage (TN) Guangzhou (CN) Lomé (TG) Montevideo (UY)
Sant Feliu de Llobregat (ES) Šempeter-Vrtojba (SI)

The Edible Cities Network focuses on Edible City Solutions (ECS), defined as Nature-Based
Solutions related to urban food production, distribution and use. ECS can include, for example,
neighbourhood gardens, bee keeping, sheep breeding, innovative distribution channels, green
facades or high-tech indoor farming services, joint cooking and eating, and provision of locally
produced food to shops and restaurants. The project shall demonstrate that ECS can make
cities healthier, greener and more enjoyable, can create new green businesses and jobs, and
can empower local communities to overcome social problems. EdiCitNet implements, monitors
and transfers ECS in close cooperation with city authorities and other local stakeholders. It
thereby aims to increase social, environmental and economic sustainability of cities.

Drawing on knowlegde from projects


funded by the European Union

Image: Oslo Living Lab - Photo © Stephanie Degenhardt


SCOPE NBS related to urban food production, distribution and use

Approach to Impact Assessment Main Challenges addressed

A long list of potential indicators for measuring


social, environmental or economic performance of 1. Climate Resilience
ECS is provided based on an extensive review of
scientific literature. Indicators that are expected 2. Water Management
to be relevant for many ECS have been included in
the EdiCitNet Toolbox (developed in another work 3. Natural and Climate Hazards
package). For more specific, detailed examples
of monitoring, the EdiCitNet impact assessment 4. Green Space Management
takes its point of departure in the ECS that are
defined in the implementation plans for the Living 5. Biodiversity
Labs in the Front-Runner Cities. Indicators are se-
lected according to the expressed goals and possi- 6. Air Quality
bly anticipated side-effects of each ECS. Methods
are chosen or adapted and agreed upon under 7. Place Regeneration
consideration of scientific soundness and human
resources locally available. Data will be stored as 8. Knowledge and Social Capacity Building
part of the online open access EdiCitNet database
developed in the EdiCitNet Toolbox. 9. Participatory Planning and Governance

10. Social Justice and Social Cohesion


Involved Stakeholders and roles
11. Health and Wellbeing
The EdiCitNet impact assessment team is centred
around the work package “Documentation and Mo- 12. New Economic Opportunities & Green Jobs
nitoring”. Members of the work package are city
administrations of Front-Runner Cities and an in-
terdisciplinary group of research partners from a
broad variety of scientific disciplines. The city ad- Lessons learned
ministrations are responsible for data collection
in practice, either through own staff or delegated Implementation of Edible City Solutions through
to other stakeholders, volunteers, students, ECS co-creation can be very time-consuming, and lo-
participants, etc. The researchers’ role is to assist cal actors do not necessarily perceive impact as-
the city administrations in selecting meaningful sessment as a priority at an early stage. Meaning-
indicators, and appropriate and feasible methods ful delimitations of ECS may vary depending upon
for data collection and storage. Experts of different local contexts, such as geography, target group,
fields need to be matched with relevant cities. The type of produce, etc. It is of utmost importance that
scientists also facilitate dialogue among cities, and assessment indicators are seen to be meaningful
ensure that comparable data is produced in different by the local actors who are responsible for data
Living Labs as far as applicable. collection. Scientists must facilitate to match in-
tended aims of an ECS with suitable indicators,
and assist stakeholders in selecting or developing
methods for data collection that are both scienti-
Municipal Administrations (FR/FL) fically sound and feasible in the light of local per-
sonnel, knowledge, time and financial resources.
Citizens Successful impact assessment further depends on
ECS coordinators and participants having access
Scientists / Academia to convenient tools for data collection, storage
and management.
NGOs
Learn more
Green businesses www.edicitnet.com

The EdiCitNet project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No. 776665.
Drawing on knowlegde from projects
funded by the European Union

URBiNAT
Urban innovative & inclusive Nature
Porto (PT) Nantes (FR) Sofia (BG) Siena (IT) Brussels (BE)
Nova Gorica (SI) Høje-Taastrup (DK) Khorramabad (IR) Shenyang (CN)

URBiNAT challenges the conventional nature-based solutions definitions by not only integ-
rating solutions inspired by nature, as the territorial and technological solutions, comprising
products and infrastructures, but also including the participatory and social and economic
solutions, comprising processes and services, that reinforce put thein dialogue between the
physical structure and the social dimension of the public space. The goal is to bring these two
plans of the public space to a living interaction, building collective awareness on commonali-
ties, both material and immaterial and, by raising the collective understanding of the human
and non-human urban dimensions, promoting the co-creation, co-development, co-implemen-
tation and co-evaluation of solutions inspired by nature and in human-nature.

Image: NBS co-selection in Corujeira Primary School, Porto


Photo © CES, Carlos Barradas
SCOPE A cluster of NBS as a healthy corridor on district level

Approach to Impact Assessment Main Challenges addressed

To analyse the effects of those implementations (and


the participation process) URBiNAT measures the
1. Climate Resilience
status quo before the participation process and after
some month of use of the new clustered NBS at dis- 2. Water Management
trict level with mixed methods. (1) A Neighbourhood
Survey asks about physical activity, social activity, 3. Natural and Climate Hazards
wellbeing, health and the satisfaction/dissatisfaction
with the environment at district level. A control group 4. Green Space Management
helps to filter out effects that are not attributable to
the implemented changes. (2) Open Spaces are ob- 5. Biodiversity
served with the technique of behavioural mapping,
while (3) sample measures capture environmental 6. Air Quality
quality. (4) Spatial GIS analysis and statistical data
complete those quantitative set of indicators. In addi- 7. Place Regeneration
tion, a number of qualitative methods like interviews,
walkthrough and photo voice enriches the understan- 8. Knowledge and Social Capacity Building
ding of the district, the people living there and accom-
panies the entire process of implementation. 9. Participatory Planning and Governance

Involved Stakeholders and roles 10. Social Justice and Social Cohesion

In URBiNAT transdisciplinary local taskforces are im- 11. Health and Wellbeing
plemented which are based on municipal administra-
tions and local universities as key stakeholders. Local 12. New Economic Opportunities & Green Jobs
participation and planning experts cover the imple-
mentation process. Through Schools, kindergartens,
NGOs and housing associations the connection to citi-
zen is established and maintained. In addition, scien- Lessons learned
tific expertise on the tackled challenges is brought in
by academic partners within the project. Finally, linka- As a highly participatory innovation action URBiNAT
ges to regional and national statistics authorities are needed to define an Impact Assessment without
established by the cities to access existing data sets. knowing what NBS the people living in those neigh-
Together these transdisciplinary group of stakeholders bourhoods would choose. Thus, the impact assess-
are the foundation for the local living labs in URBiNAT ment strategy focused on the healthy corridor as
and ensure a flow of data for monitoring and impact a cluster of NBS which opened the perspective to
evaluation. This data and it‘s analysis is shared within assess the benefits for the whole district. To trans-
the project and beyond via the observatory platform form the districts into study areas which ensure an
(www.urbinatobservatory.eu). efficient flow of necessary data several hurdles have
to be taken. The transdisciplinary team has to over-
Municipal Administrations come barriers in language and knowledge between
stakeholders. It is important to come to a common
understanding and agreement on the effects of inte-
Regional/national statistics authority
rest and a realistic timing when they will occur. There
are several effects that will not be immediately visib-
Citizen le, thus the differences between short-, middle- and
long-term effects need to be taken carefully into ac-
Planning experts count. It is therefore essential to allocate a realistic
amount of time and resources to set up the team itself
Scientists / Academia as well as to develop and conduct the impact assess-
ment strategy before and after the implementation
NGOs phase. In addition, it is important to underline the role
of participatory activities which can give a perception
of change and immediate benefits for the community.
Schools and kindergartens
Learn more
Housing associations www.urbinat.eu

The URBiNAT project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 776783
4 INDICATORS OF NBS PERFORMANCE AND IMPACT
Coordinating Lead authors
Wendling, L., Dumitru, A.

Lead authors
Arnbjerg-Nielsen, K., Baldacchini, C., Connop, S., Dubovik, M., Fermoso, J., Hölscher, K.,
Nadim, F., Pilla, F., Renaud, F., Rhodes, M. L., San José, E., Sánchez, R., Skodra, J., Tacnet,
J.-M., Zulian, G.

Contributing authors
Allaert, K., Almassy, D., Ascenso, A., Babí Almenar, J., Basco, L., Beaujouan, V., Benoit, G.,
Bockarjova, M., Bode, N., Bonelli, S., Bouzouidja, R., Butlin, T., Calatrava, J., Calfapietra, C.,
Cannavo, P., Capobianco, V., Caroppi, G., Ceccherini, G., Chancibault, K., Cioffi, M., Coelho,
S., Dadvand, P., de Bellis, Y., de Keijzer, C., de la Hera, A., De Vreese, R., Decker, S.,
Djordjevic, S., Dowling, C., Dushkova, D., Eiter, S., Faneca, M., Fatima, Z., Ferracini, C.,
Fjellstad, W., Fleury, G., Freyer, B., García, I., García‐Alcaraz, M., Gerundo, C., Gil-Roldán, E.,
Giordano, R., Giugni, M., Goličnik Marušić, B., Gómez, S., González, M., Gonzalez-Ollauri, A.,
Guidolotti, G., Haase, D., Heredida, J., Hermawan, T., Herranz-Pascual, K., Jermakka, J.,
Jones, L., Kiss, M., Kraus, F., Körmöndi, B., Laikari, A., Laille, P., Lemée, C., Llorente, M.,
Lodder, M., Macsinga, I., Maes, J., Maia, S., Manderscheid, M., Manzano, M., Martelli, F.,
Martins, R., Mayor, B., McKnight, U., Mendizabal, M., Mendonça, R., Mickovski, S.B., Miranda,
A.I., Moniz, G.C., Munro, K., Nash, C., Nolan, P., Oen, A., Olsson, P., Olver, C., Ozturk, E.D.,
Paradiso, F., Petucco, C., Pisani, N., Piton, G., Pugliese, F., Rasmussen, M., Ravknikar, Ž,
Reich, E., Reichborn-Kjennerud, K., Rinta-Hiiro, V., Robles, V., Rodriguez, F., Roebeling, P.,
Ruangpan, L., Rugani, B., Rödl, A., Sánchez, I., Sánchez Torres, A., Sanesi, G., Sanz, J.M.,
Scharf, B., Silvestri, F., Spano, G., Stanganelli, M., Szkordilisz, F., Tomé-Lourido, D., Vay, L.,
Vela, S., Vercelli, M., Villazán, A., Vojinovic, Z., Werner, A., Wheeler, B., Young, C., Zorita, S.,
Zandersen, M., zu-Castell Rüdenhausen, M.

115
Summary

What is this chapter about?

This chapter introduces 12 categories of societal challenges that NBS can address
(Section 4.1). These are conceptually mapped against the UN Sustainable
Development Goals. For each of the 12 societal challenge areas, Section 4.2
outlines and lists indicators to evaluate the performance and impact of NBS. It
reviews the different types of NBS, gives examples of each NBS type, and lists
the indicators related to the particular societal challenge in a series of tables.
Associated methodologies are compiled in the related Appendix of Methods. To
help navigate, the indicators are classified as structural, process-based or
outcome-oriented. Structural indicators are particularly useful during the NBS
planning process and can help identify where resources may be lacking or
highlight policy and/or procedural gaps that require attention. Process-based
indicators can provide information about the value or impacts of the collaborative
processes that underpin NBS (co-creation, co-implementation and co-
management). The outcome-oriented indicators are useful to understand NBS
performance by establishing an understanding of baseline (pre-NBS) conditions
and following changes to these conditions after NBS implementation. We
distinguish between recommended and additional indicators. Recommended
indicators are considered the most important ones to monitor NBS impact.
Additional indicators can provide highly valuable information, depending on local
context and particular data needs. The chapter concludes with a reflection on the
importance of critical thinking to select the right indicators for a holistic
assessment of NBS and the development of emerging indicators (Section 4.3).

How can I use this chapter in my work with NBS?

This chapter helps to select the most appropriate indicators to assess the
performance and impact of a given NBS. As resources are limited and it is simply
not possible to monitor every single indicator, this buffet-style approach enables
tailoring of a monitoring programme to address a specific context, both with
respect to the challenges addressed and the NBS implemented in response.

When should I use this knowledge in my work with NBS?

Selection of indicators can occur at any time during the cycle of adaptive
management of NBS. The initial monitoring and assessment plan identifies “must-
have” outcomes that can be linked to specific indicators. For example, if the
primary objective of a given NBS is to attenuate flooding then indicators related
to the impacts of floods (extent of flooded land, duration of flooding, number of
buildings and/or persons affected, etc.) are critical to evaluate NBS impact.
During the NBS co-creation process, review of planned NBS impact indicators can
help to identify potential additional benefits and inform NBS design. Indicators
can be added or replaced at any time in response to observed changes or new
challenges (adaptive monitoring).

116
How does this chapter link with the other parts of the handbook?

The previous chapters have detailed the concept of NBS and briefly described
how NBS can support relevant public policies, why it is important to monitor NBS
performance and evaluate their impacts, and how to develop a monitoring and
evaluation strategy. This chapter focuses on which indicators to use in different
local contexts in order to understand NBS performance and impacts. Chapter 4
should be read in conjunction with the Appendix of Methods, where the specific
details of each indicator are further clarified, along with a brief methodology. The
following Chapters 5 and 6 expand upon the list of indicators presented here by
illustrating the application of selected indicators to NBS in different contexts,
including NBS specifically designed for disaster risk reduction (DRR). Chapter 7
describes the different types of NBS monitoring data and provides detailed
information about how to acquire and evaluate the quality these data.

4.1 Societal challenge areas addressed by NBS

The 2017 EKLIPSE Expert Working Group impact evaluation framework report
(Raymond et al., 2017) identified ten challenge areas related to climate resilience
in urban areas. The present report expands these original ten challenge areas to
12 separate societal challenge areas that can potentially be addressed by NBS
(Figure 4-1). In addition to presenting a suite of indicators applicable to each
challenge area, methods of indicator determination are presented in the separate
report Evaluating the Impact of Nature-based Solutions: Appendix of Methods to
support the application of impact indicators. The overarching objective of this
Handbook and the accompanying Appendix of Methods is to provide standardized
guidance and methods of indicator determination to support establishment of a
robust European evidence base on NBS performance and impact. In order to
compare different types of NBS, implemented in different environments and at
varying scale we need to measure the same variables, using the same methods
and report these outcomes using the same units of measure.

The 12 challenge areas elaborated herein are:

1. Climate Resilience
2. Water Management
3. Natural and Climate Hazards
4. Green Space Management
5. Biodiversity Enhancement
6. Air Quality
7. Place Regeneration
8. Knowledge and Social Capacity Building for Sustainable Urban Transformation
9. Participatory Planning and Governance
10. Social Justice and Social Cohesion
11. Health and Wellbeing
12. New Economic Opportunities and Green Jobs

117
Figure 4-1. Conceptual mapping of societal challenge areas that can be addressed by NBS onto the triad of
People, Planet, Prosperity pillars of sustainable development

Climate Resilience: Nature-based solutions are capable of providing resilience


to the impacts of climate change through the provision of ecosystem services,
and by enhancing social awareness and actions to combat climate change. The
co-benefits delivered by NBS support climate change mitigation and adaptation
efforts, particularly in urban areas, contributing to the liveability of cities.

Water Management: Nature-based solutions provide an excellent opportunity


to address a diversity of issues associated with anthropogenic impacts on the
water cycle. These include poor water quality, water availability for extraction,
groundwater and surface water levels, recharging of aquifers, stormwater
management, water treatment, wetland habitat management, soil water
management, and ecological quality.

Natural and Climate Hazards: Risk is a combination of hazard and (negative)


consequences. Nature-based solutions employed for disaster risk reduction are
expected to reduce risk level (i.e., influence risk components corresponding to
hazard or vulnerability). At the same time, NBS deliver further social, human,
and environmental co-benefits. This challenge category was expanded based
upon the further development of the “Coastal Resilience” challenge area
described in the EKLIPSE Expert Working Group impact evaluation framework
(Raymond et al., 2017) to include a wider array of climate-related and natural
hazards.

118
Green Space Management: Green space management refers to the planning,
establishment and maintenance of green and blue infrastructure in urban areas.
Green and blue infrastructure (abbreviated as urban green infrastructure, UGI)
are a type of NBS that refers specifically to the strategically managed network of
natural and semi-natural ecosystems within urban boundaries. UGI provides a
range of ecological and socio-economic benefits (Raymond et al., 2017) and, if
correctly managed, contributes to solutions for numerous challenges such as air
and noise pollution, heat waves, flooding and concerns regarding public well-
being (Maes et al., 2019). NBS support the wider deployment of green and blue
infrastructure (EC, 2019a; EC, 2019b), thus supporting the EU Green
Infrastructure Strategy (EC, 2013) and the EU Biodiversity Strategy for 2030 (EC,
2020).

Biodiversity Enhancement: Biodiversity loss and ecosystem collapse are


among the greatest threats society faces in the near term. There are five primary
direct drivers of biodiversity loss: changes in land and sea use, overexploitation,
climate change, pollution, and invasive alien species. The link between climate
change and biodiversity loss involves a feedback loop whereby climate change
accelerates loss of natural capital, which is in turn a key driver of climate change.
NBS support the EU Biodiversity Strategy for 2030 (EC, 2020) through the
purposeful establishment of protected areas and restoration of degraded
ecosystems. The enhancement and/or conservation of biodiversity was
considered as part of the Green Space Management challenge in the EKLIPSE
Expert Working Group impact evaluation framework (Raymond et al., 2017).
Here, we consider Biodiversity Enhancement as a separate challenge area.

Air Quality: NBS based on the creation, enhancement, or restoration of


ecosystems in human-dominated environments play a relevant role in removing
air pollutants and carbon dioxide, reducing the air temperature (which slows
down the creation of secondary pollutants) and increasing oxygen concentration,
contributing to a beneficial atmospheric composition for human life.

Place Regeneration: Urbanisation has a lasting impact on the natural


environment of towns and cities, not only visible through dereliction, but also
through increasing environmental footprint fuelled by economic growth and
unsustainable patterns of consumption. Nature-based solutions hold the potential
to contribute to the aim of ensuring successful achievement of sustainable place
regeneration by way of enhancing the green space and people-nature connection,
as well as using fewer environmental resources, enhancing place resilience to
natural disasters, fostering collective participation and social cohesion, and
improving individual wellbeing (Korkmaz and Balaban, 2020; Roberts and Sykes,
2000; Xiang et al., 2017).

Knowledge and Social Capacity Building for Sustainable Urban


Transformation: Sustainable urban transformation delineates sustainable urban
structures and environments, as well as radical social, economic, cultural,
organizational, governmental, and physical change processes (Ernst et al., 2016;
McCormick et al., 2013). Knowledge and social capacity building through
educational initiatives can contribute to the complex enterprise of amassing
resources for sustainable urban places. This challenge area is a new addition to

119
the original ten challenges described in the EKLIPSE Expert Working Group impact
evaluation framework (Raymond et al., 2017).

Participatory Planning and Governance: Nature-based solutions demand


approaches to planning and governance frameworks that support accessibility to
green spaces, while maintaining their quality for ecosystem services provision.
Urban environmental transformation is a highly complex undertaking that
requires open collaborative governance and robust capacities for participatory
planning. Nature-based solutions already implemented and functional across
Europe have contributed a wealth of knowledge in the area of participatory
planning and governance, indicating, for instance, that successful outcomes call
for openness to learning and experimenting along other urban actors so as to co-
create and co-maintain nature-based solutions while shaping institutional spaces
in cities that allow for this co-creation, social innovation and collaboration to
continue (Frantzeskaki, 2019). Significantly, open collaborative governance and
participatory planning invested in nature-based solution strategies bring forward
opportunities for social transformation and increased social inclusiveness in cities
(Wendling et al., 2018).

Social Justice and Social Cohesion: Nature-based solutions have been linked
to the notion of environmental justice across studies that explore the role of
supporting urban processes involving equal access to neighbourhood green space
in fostering social cohesion (e.g., bridging and bonding social capital) towards the
cultural integration of typically-excluded social groups, like elderly, immigrants,
persons with disabilities, etc. (i.e., recognition-based justice) (Ibes, 2015; Kweon
et al., 1998; Raymond et al., 2017; Raymond et al., 2016; van Den Berg et al.,
2017). Recently, Gentin et al. (2019) analysed the premises for a nature-based
integration of immigrants in Europe and urged on researchers to set aside
descriptions and analyses of immigrants’ perceptions or use of nature, and turn
their focus towards exploring and developing nature-based solutions for the
purposes of social integration.

Health and Wellbeing: Critical social and environmental determinants of health,


including clean air, safe drinking water, sufficient food and secure shelter, are
impacted by climate change 44. More than half of the world’s population lives in
urban areas (towns and cities), and this number is projected to increase to two
in three people by 2050 45. Climate change and other environmental issues affect
all categories of population, however it is most threatening in urban areas where
the majority of the population live. This means that the consequences of climate
change, poor air quality and other current concerns are often very obvious and
disruptive to urban living, and can affect services such as sanitation leading to
public health issues.

New Economic Opportunities and Green Jobs: Key criteria of NBS are their
cost-effectiveness, and their capacity to simultaneously provide environmental,
social and economic benefits in support of resilience building. The adoption and
implementation of NBS has the potential to create new economic opportunities

44
https://www.who.int/news-room/fact-sheets/detail/climate-change-and-health
45
http://www.un.org/en/development/desa/news/population/world-urbanization-prospects-2014.html

120
and jobs in the green sector by enabling low-carbon, resource-efficient and
socially inclusive economic growth. Within this paradigm, economic growth is
driven by public and private investment in activities, infrastructure and assets
that support reduced emissions of carbon and pollutants, and increased energy
and resource efficiency whilst enhancing biodiversity and the provision of
ecosystem services.

4.2 Recommended and Additional indicators for NBS impact


assessment

The NBS impact evaluation relies strongly on the adoption of quantitative and
qualitative impact markers – the performance and impact indicators. These serve
as means for assessing the progress of an adopted pathway targeted at achieving
specific objectives, including those of various temporal and spatial scales. The
Recommended indicators for each of the twelve societal challenge areas
presented herein serve as a ‘starting point’ for evaluating the NBS impact, and
they are considered as the primary indicators to be addressed when creating NBS
monitoring and evaluation schemes. The Recommended indicators listed herein
represent a foundation of performance and impact indicators to be considered for
all NBS projects and that they should also provide sufficient flexibility to be
applicable to all NBS scenarios.
The list of Additional indicators comprise the remaining NBS performance and
impact indicators adopted by the H2020 NBS project teams involved in the
production of this Handbook (see Chapter 1), and can be used to complement
the list of Recommended indicators for a more holistic assessment. The selection
of Additional indicators aligns with specific NBS project objectives. Some
examples of Additional indicator selection are presented in the following chapter
(Chapter 5).
A suite of Recommended and Additional indicators for each of the twelve
identified societal challenge areas are outlined in the following sub-sections.
Indicators of NBS impact have been classified as structural, process or outcome
based (Donabedian, 1966) to support the selection of a suite of indicators that
holistically address the process of NBS co-creation, co-implementation and co-
management.
• Structural indicators (S) – refer to supporting infrastructure and
resources in place to achieve the desired goals (people, material, policies
and procedures)

• Process indicators (P) – refer to the efficiency, quality, or consistency


of specific procedures employed to achieve the desired goals

• Outcome indicators (O) – refer to accomplishments or impacts

Whilst this classification does not explicitly refer to the timing of indicator use, it
follows that the structural indicators may be most useful during the planning of
NBS, i.e., to determine what resources or supporting policies may be needed to
ensure the success of the proposed NBS action. The process indicators are useful

121
to evaluate the methods used to co-create, co-implement and co-manage NBS,
and so can be applied throughout the adaptive management cycle but are most
relevant during periods of intense activity. A large proportion of the NBS impact
indicators listed herein are primarily focused on the impact or end result of NBS
actions.

Note that nearly all of the indicators listed here can be used prior to NBS
implementation to establish an understanding of pre-NBS, or ‘baseline’,
conditions as well as during and following NBS actions. Comparison of pre-NBS
measures with additional measurements during or following NBS implementation
will show how conditions change with time. Measurements collected over time
can be used to illustrate the longer-term impacts of NBS and how different
outcomes are realised with time. It is important to be careful interpreting data,
as not all observed changes can necessarily be directly attributed to NBS actions.
In some cases the impacts of NBS may be more clear when comparing
measurements taken at the same time at two different sites, i.e., the NBS site
and an analogous location without NBS (a ‘control site’). This is particularly
important when there are multiple changes to an area or there are external
influences on the system, such as significant changes to hydrologic regime from
the original ‘baseline’ condition.

The following tables also show the applicability of each indicator to different types
of NBS. Nature-based solutions can be broadly grouped based upon their primary
objective or function and by the level of ecosystem intervention. The following
NBS typology proposed by Eggermont et al. (2015) has been widely adopted
(Figure 4-2):

• Type 1 NBS – minimal or no intervention in ecosystems, with objectives


related to maintaining or improving delivery of ecosystem services within
and beyond the protected ecosystems

• Type 2 NBS – extensive or intensive management approaches seeking


to develop sustainable, multifunctional ecosystems and landscapes in
order to improve delivery of ecosystem services relative to conventional
interventions

• Type 3 NBS – characterised by highly intensive ecosystem management


or creation of new ecosystems

122
Figure 4-2. Schematic representation of NBS typology (adapted from Eggermont et al., 2015)

Type 1 NBS include protection and conservation strategies, urban planning


strategies, and (environmental) monitoring strategies. Due to their nature, Type
1 NBS fall largely within the domain of governance, with implementation of Type
1 NBS strategies potentially limited or driven by a range of biophysical, social and
institutional factors. Type 2 NBS are comprised of various sustainable
management practices. Type 3 NBS are newly-created ecosystems, and therefore
are the most “visible” solutions. Examples of Types 1-3 NBS may include
(Cohen-Shacham et al., 2016; Eggermont et al., 2015; EC, 2015; Somarakis et
al., 2019):

Type 1 NBS
• Protection and conservation strategies
 Establishment of protected areas or conservation zones
 Limitation or prevention of specific land use and/or practices
 Ensuring of continuity of ecological networks (protection from
fragmentation)
 Maintenance or enhancement of natural wetlands

• Urban planning strategies


 Ensuring of continuity of ecological network
 Controlling urban expansion

• Monitoring
 Regular monitoring of physical, chemical or biological indicators
Type 2 NBS
• Sustainable management protocols
 Integrated pest/weed management
 Spatial and/or time and frequency aspects of integrated and ecological
management plans
 Creation and preservation of habitats and shelters to support
biodiversity (e.g., insect hotels for wild bees, next boxes for native bats
and birds, stopover habitat/”rest stops” for migratory birds)

123
 Installation of apiaries
 Sustainable fertiliser use
 Control of erosion through management of grazing animal stocking
density and exclusion of grazing animals from riparian areas
 Composting of organic wastes and reuse of composted material
 Integrated water resource management
 Protection of plant resources from pest and disease
 Aquifer protection from pollution and sustainable management of
withdrawals
Type 3 NBS
• Green space - multifunctional open space characterised by natural vegetation
and permeable surfaces
 Urban parks and gardens of all sizes
 Heritage park
 Botanical garden
 Community garden
 Cemetery
 Schoolyards and sports fields
 Meadow
 Green strips
 Green transport track
 “Multifunctional” dry detention pond or vegetated drainage basin

• Trees and shrubs


 Forests (including afforestation)
 Orchards
 Vineyards
 Hedges/shrubs/green fences
 Street trees

• Soil conservation and quality management


 Slope revegetation
 Cover crops
 Windbreaks
 Conservation tillage practices
 Permaculture
 Deep-rooted perennials
 Organic matter enrichment (manure, biosolids, green manure, compost,
etc.)
 Inorganic soil conditioners and amendments (biochar, vermiculite, etc.)

• Blue-green space establishment or restoration


 Riparian buffer zones
 Mangroves
 Saltmarsh/seagrass
 Intertidal habitats
 Dune structures

• Green built environment


 Green roof

124
 Green-blue roof
 Green wall/façade
 Green alley
 Infiltration planters and tree boxes
 Temporary and/or small-scale interventions including green furniture,
green living rooms, etc.

• Natural or semi-natural water storage and transport structures


 Surface wetland
 Floodplains, floodplain reconnection with rivers
 Restoration of degraded waterbodies
 Restoration of degraded waterways, including re-meandering of
streams and river daylighting
 Retention pond/wet detention pond

• Infiltration, filtration, and biofiltration structures


 Infiltration basin
 Vegetated filter strip
 Rain garden
 Wet/dry vegetated swale, with or without check dams
 Subsurface wetland or filtration system
 Bioretention basin/bioretention cell

The preceding list of NBS is non-exhaustive and is intended only to provide


examples of different types of NBS per the Type 1-3 classification system. The
tables in this chapter indicate in general whether a particular indicator is
applicable to Type 1, 2 or 3 NBS; however, the wide variety of NBS actions make
consideration of all possible combinations of NBS and indicator application quite
challenging. The NBS type 1-3 indicator applicability shown in the following tables
should be considered a guide.

4.2.1 Climate Resilience

Indicators in the Climate Resilience challenge area primarily address:

• Direct impacts of NBS on greenhouse gas emissions via carbon storage and
sequestration in vegetation and soil;

• Indirect impacts of NBS on avoided greenhouse gas emissions from various


activities, through the provision of passive cooling, insulating and/or water
treatment; and,

• Impacts of NBS on temperature and human comfort

Primary among the Recommended indicators for the Climate Resilience challenge area
is carbon sequestration. Accounting for C stored in soil and vegetation, particularly in
an urban area, can provide a tangible evaluation of local climate change mitigation and

125
the impacts of local land use, planning and decision-making. This is reflected by the
total quantity of carbon removed or stored in soil and vegetation (indicator 1.1) as it
provides a measure for direct carbon sequestration by NBS. In contrast, the quantity
of avoided greenhouse gas emissions due to reduced building consumption (indicator
1.2) reflects the cooling and/or insulating capacity of NBS, resulting in lesser energy
use for building cooling or heating.

Nature-based solutions can be an effective means to combat urban heat islands.


Although NBS cannot alter the weather, the presence of (large-scale) NBS may provide
sufficient cooling to locally mitigate high temperatures during heat wave events. NBS
can support reduced energy use and improved thermal comfort by moderating the
urban microclimate (Demuzere et al., 2014), which is reflected by monthly mean daily
maximum (TXx, indicator 1.3) and minimum (TNn, indicator 1.4) temperature, which
provide a measure of the local cooling or warming effect of NBS. These indicators are
related both to building energy use as well as human comfort. Indicator 1.5, heatwave
incidence, reflects prolonged periods of abnormally high temperatures, and can be
used to measure the local impact of NBS on ambient temperatures during these
periods,

Additional indicators are listed that can be employed to quantify specific parameters
generally related to NBS-provided ecosystem services in support of climate resilience.
They can further be utilised to complement the assessment of the Recommended
indicators for generating a more holistic picture of the local NBS performance.

Table 4-1. Indicators related to Climate Resilience classified as structural (S), process focused (P) or
outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Total carbon
removed or stored in
1.1 vegetation and soil kg/ha/y O ● ● ●
per unit area per
unit time

Avoided greenhouse
gas emissions from
1.2
reduced building
t CO2e/y O ● ●
energy consumption

Monthly mean value


1.3 of daily maximum °C O ● ●
temperature (TXx)

Monthly mean value


1.4 of daily minimum °C O ● ●
temperature (TNn)

126
Heatwave incidence:
Days with
1.5
temperature >90th
No./y O ● ●
percentile, TX90p

ADDITIONAL

2.1.1 Total carbon stored


2.1.2 in vegetation
kg/ha/y O ● ● ●

2.1.3 Total leaf area m2 O ● ● ●

Carbon storage
2.1.4
score
kg/day O ● ● ●

2.1.5
2.1.6
Soil carbon content ton/ha O ● ● ●

Rate of soil carbon


2.1.7
decomposition
% p.a. O ● ● ●

Energy use savings


2.2 due to NBS kWh/y O ● ●
implementation

Carbon emissions
2.3 due to building t CO2e/y O ●
cooling

Carbon emissions
due to treatment of
2.4
runoff water
t CO2e/y O ● ● ●
(combined sewers)

2.5 Soil temperature °C O ● ● ●

Total surface area of


2.6
wetlands
ha O ● ● ●

Surface area of
2.7 restored and/or ha O ● ● ●
created wetlands

Aboveground tree
2.8
biomass
t/ha O ● ● ●

Human comfort:
2.9.1 Universal Thermal °C O ● ●
Climate Index

Thermal Comfort
2.9.2
Score
unitless O ● ●

Human comfort:
Physiological
2.9.3
Equivalent
°C O ● ●
Temperature

127
Mean or peak
daytime temperature
– Predicted Mean
2.9.4
Vote-Predicted
unitless O ● ●
Percentage
Dissatisfied

Urban Heat Island


2.10.1
(incidence)
°C O ● ●

Number of combined
2.10.2 tropical nights and No. O ● ●
hot days

Thermal Storage
2.10.3
Score
J O ● ●

2.10.4 Thermal Load Score °C O ● ●

Peak summer
2.11 temperature (GI- °C O ● ●
Val)

Maximum surface
2.12
cooling
°C O ● ●

2.13.1 Mean local daytime


2.13.2 temperature
°C O ● ●

2.13.1 Peak local daytime


2.13.2 temperature
°C O ● ●

Daily temperature
2.14
range
°C O ● ●

2.15
2.15.1 Air cooling °C O ● ●
2.15.2

Tree shade for local


2.16
heat reduction
m2 O ● ● ●

Rate of
2.17
evapotranspiration
mm/day O ● ● ●

Land surface
2.18
temperature
°C O ● ● ●

Surface reflectance -
2.19
albedo
unitless O ● ●

Carbon emissions
2.20
from vehicle traffic
t C/y O ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving delivery
of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

128
4.2.2 Water Management

The diversity of potential benefits, co-benefits, and trade-offs related to NBS use
for water management is reflected in the comprehensive list of Recommended
indicators presented. These Recommended indicators were selected by members
of a range of EU H2020 NBS projects working across urban, peri-urban, and rural
areas. The Recommended list is representative of this diversity of approaches.

From the comprehensive list of Water Management indicators proposed by the


H2020 NBS project teams, the list of Recommended Indicators was selected
based on those that were considered to be the key drivers of nature-based
solution implementation, and thus those that were relevant to the highest
proportion of nature-based solution initiatives. The indicators selected as
Recommended address the potential benefits, co-benefits, and trade-offs
associated with changes to surface water runoff volume (3.1) and to water quality
(3.2-3.6).

The Additional indicators address a wide range of applicable metrics for the
assessment of NBS impact from a broad perspective, further exploring potential
impacts on soil-water interactions, additional aspects of stormwater and excess
runoff management, and actions pertinent to the implementation of the Water
Framework Directive 46, including quantitative, hydromorphological, ecological
and physico-chemical status of surface and groundwaters.

Table 4-2. Indicators related to Water Management classified as structural (S), process focused (P) or
outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Surface runoff in
3.1 relation to mm/% O ● ● ●
precipitation quantity

Water quality:
3.2
general urban
various O ● ● ●

Water quality: TSS


3.3
content
mg/L O ● ● ●

Nitrogen and
3.4 phosphorus % O ● ● ●
concentration or load

46
Directive 2000/60/EC, OJ L 327, 22.12.2000

129
Metal concentration
3.5
or load
% O ● ● ●

Water quality: total


faecal coliform
3.6
bacteria content of
No. O ● ● ●
NBS effluents

ADDITIONAL

4.1 % or
4.2
Infiltration rate
mm/h
O ● ● ●

4.1
4.2
Infiltration capacity mm/d O ● ● ●

Rate of mm/m2
4.3
evapotranspiration day
O ● ● ●

4.4 Peak flow variation % O ● ● ●

4.5 Flood peak reduction % O ● ● ●

4.5 Flood peak delay h O ● ● ●

4.6 Height of flood peak m3/s O ● ● ●

4.6 Time to flood peak h O ● ● ●

4.7 Flood Excess Volume m3 O ● ● ●

Rainfall interception
4.8
of NBS
mm/h O ● ● ●

Runoff rate for


4.9 different rainfall m3/s O ● ● ●
events

4.10 Run-Off Score (ROS) unitless O ● ● ●

Rainfall storage
4.11
capacity of NBS
mm/% O ● ● ●

Quantitative status Good or


4.12
of groundwater Poor
O ● ● ●

Depth to
4.13
groundwater
m O ● ● ●

Chemical status of Good or


4.14
groundwater Poor
O ● ● ●

130
Trend in piezometric
4.15
levels
m3/y O ● ● ●

Groundwater
4.16
Exploitation Index
% O ● ● ●

Aquifer surface ratio


4.17
with excessive nitrate
% O ● ● ●

Aquifer surface ratio


4.18 with excessive % O ● ● ●
arsenic

Rainwater or
4.19 greywater use for m3/y O ● ● ●
irrigation purposes

Water Exploitation
4.20
Index
% O ● ● ●

Water dependency
4.21
for food production
m3 O ● ● ●

Calculated drinking
4.22
water provision
m3/ha/y O ● ● ●

Net surface water


4.23
availability
m3/y O ● ● ●

Volume of water
removed from
4.24
wastewater
m3/y O ● ● ●
treatment system

Volume of water
slowed down
4.25
entering sewer
m3/s O ● ● ●
system

Total surface area of


4.26
wetlands
ha O ● ● ●

Surface area of
4.27 restored and/or ha O ● ●
created wetlands

4.28 Soil water saturation % O ● ● ●

Soil water retention


4.29
capacity
m3/m3 O ● ● ●

4.30 Stemflow rate mm/h O ● ● ●

Percolation rate
4.31 under different mm/d O ● ● ●
rainfall events

131
Dissolved oxygen
4.32 content of NBS mg/L O ● ● ●
effluents

4.33 Eutrophication unitless O ● ● ●

4.34 pH of NBS effluents unitless O ● ● ●

Electrical
4.35 conductivity of NBS µS/cm O ● ● ●
effluents

High,
Physico-chemical
Good,
4.36 quality of surface
Moderate,
O ● ● ●
waters
Poor, Bad

Total pollutant
4.37 discharge to local unitless O ● ● ●
waterbodies

Water quality: basic


4.38
physical parameters
various O ● ● ●

Total PAH content of


4.39
NBS effluents
ng/L O ● ● ●

Total organic carbon


4.40 content of NBS mg/L C O ● ● ●
effluents

High,
General ecological
Good,
4.41 status of surface
Moderate,
O ● ● ●
waters
Poor, Bad

Ecological potential Maximum,


for heavily modified Good,
4.42
or artificial water Moderate,
O ● ● ●
bodies Poor, Bad

High,
Biological quality of Good,
4.43
surface waters Moderate,
O ● ● ●
Poor, Bad

Extended Biotic
Index: total number
4.44 and species richness unitless O ● ● ●
of aquatic
macroinvertebrates

Morphological
4.45
Quality Index
unitless O ● ● ●

High,
Hydromorphological
Good,
4.46 quality of surface
Moderate,
O ● ● ●
waters
Poor, Bad

132
Fluvial Functionality
4.47
Index
unitless O ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.3 Natural and Climate Hazards

Indicators of NBS impact with respect to natural and climate hazards provided in
this list are expected to be useful to measure the effectiveness of NBS. Application
of these indicators will enable measurement of the effects of NBS on risk due to
natural and climatic hazards (reduction of risk, effect on one risk component).
Recommended indicators relate to three main categories and correspond to
several levels of integration ranging from global policy objectives to hazard
specific indicators.

Recommended indicators are more integrated and can be used to assess NBS
effectiveness:

• Global policy (5.1, 5.2): These integrated indicators correspond to the


way risk perception/culture is affected by the measure. Indicator 5.1 is
itself the result of a lengthy assessment process and aggregation of
several criteria.

• Vulnerability (5.3, 5.4, 5.5)

• Hazard and threat (5.6)

Additional indicators are mainly basic, unitary indicators primarily related to


hazard intensity. They are broadly listed by types of hazard (e.g., floods, coastal
erosion, landslides, water availability, and heat waves). It should be noted that
this list is non-exhaustive; however, the indicators provided herein can provide
the basis for a comprehensive NBS performance and impact monitoring scheme
focused on evaluating NBS with respect to disaster risk.

133
Table 4-3. Indicators related to Natural and Climate Hazards classified as structural (S), process focused
(P) or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

5.1 Disaster Resilience unitless S ● ● ●

Disaster-risk
5.2 informed unitless S ●
development

Mean annual direct


and indirect losses
5.3
due to natural and
€ O ● ● ●
climate hazards

Risk to critical urban


5.4
infrastructure
% O ● ● ●

Number of people
adversely affected
5.5
by natural disasters
unitless O ● ● ●
each year

Multi-hazard early
5.6
warning
unitless S ●

ADDITIONAL

Urban/residential
6.1.1 areas exposed to ha O ● ● ●
risks

Productive areas
6.1.2
exposed to risks
ha O ● ● ●

Natural Areas, Site


of Community
Importance (SCI),
6.2
Special Protection
ha O ●
Areas (SPA)
exposed to risks

Inhabitants exposed
6.3.1
to risks
No./ha O ● ● ●

Area exposed to
6.3.2
flood risk
ha O ● ● ●

Local population
6.3.2 exposed to flood No./ha O ● ● ●
risk

134
Other people
(workers, tourists,
6.3.3
homes) exposed to
No./ha O ● ● ●
risk

Elderly, children,
6.3.4 disabled exposed to No./ha O ● ● ●
risk

6.4 Population
6.4.1 vulnerable to risks
No./ha O ● ● ●

Housing potentially
6.5.1
exposed to risks
No. O ● ● ●

Agricultural and
industrial buildings
6.5.2
potentially exposed
No. O ● ● ●
to risks

Strategic buildings
6.5.3
exposed to risk
No. O ● ● ●

Roads exposed to
6.6.1
risk
m/km2 O ● ● ●

Railways exposed to
6.6.2
risk
m/km2 O ● ● ●

Lifelines exposed to
6.6.3
risk
m/km2 O ● ● ●

Buildings vulnerable
6.7.1
to risks
No./km2 O ● ● ●

Transportation
infrastructure and
6.7.2
lifelines vulnerable
m/km2 O ● ● ●
to risks

Insurance against
6.8
catastrophic events
% P ●

6.9 Flood hazard unitless O ● ● ●

6.10 Flooded area ha O ● ● ●

6.11 Height of flood peak m3/s O ● ● ●

6.11 Time to flood peak h O ● ● ●

6.12 Peak flow rate m3/s O ● ● ●

6.13 Peak flood volume m3 O ● ● ●

135
6.14 Flood Excess Volume m3 O ● ● ●

6.15 Moisture Index unitless O ● ● ●

6.16 Flammability Index unitless O ● ● ●

unitless,
6.17 Soil type
qualitative
S ●

6.18 Soil shear strength kPa S ●

6.18 Soil cohesion kPa S ●

6.19 Soil temperature °C O ● ● ●

m below
Level of
6.20
groundwater table
ground O ● ● ●
surface

Slope stability factor


6.21
of safety
unitless O ● ● ●

Landslide safety
6.22
factor
unitless O ● ● ●

Landslide risk – unitless;


6.23 history of instability binominal S ●
on site (yes/no)

Occurred landslide
6.24
area
% S ●

6.25 Landslide risk % O ● ● ●

6.26
Soil mass
kg/ha O ● ● ●
movement

Velocity of occurred
6.27
landslide
m/s O ●

6.28 Erosion risk m3/year O ● ● ●

6.29
Total predicted soil
t/ha/y O ● ● ●
loss

Days with
6.30 temperature >90th % O ● ● ●
percentile, TX90p

Warm Spell
6.31
Duration Index
unitless O ● ● ●

136
6.32 Heatwave incidence No./y O ● ● ●

Human comfort:
6.33 Universal Thermal °C O ● ● ●
Climate Index

Human comfort:
Physiological
6.34
Equivalent
°C O ● ● ●
Temperature

Mean or peak
daytime
temperature –
6.35 Predicted Mean unitless O ● ● ●
Vote-Predicted
Percentage
Dissatisfied

Urban Heat Island


6.36
(incidence)
°C O ● ● ●

Effective Drought
6.37
Index
unitless O ● ● ●

Standardised
6.38
Precipitation Index
unitless S ●

Quantitative status
6.39
of groundwater
Good or Poor O ● ● ●

Trend in piezometric
6.40
levels
m3/y O ● ● ●

Groundwater
6.41
exploitation index
% O ● ● ●

Calculated drinking
6.42
water provision
m3/ha/y O ● ● ●

Water Exploitation
6.43
Index
% O ● ● ●

Net surface water


6.44
availability
m3/y O ● ● ●

Rainwater or
6.45 greywater use for m3/y O ● ● ●
irrigation purposes

Avalanche risk:
6.46
Snow cover map
unitless S ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

137
4.2.4 Green Space Management

The management of UGI interventions has impact at a range of scales, from


building and street level to district, urban, regional, national and transnational
level. Green spaces, or UGI, are a key component of many urban planning and
climate change adaptation and mitigation strategies. Related actions are included
in several transnational initiatives including, for example, the EU Strategy on Green
Infrastructure and the EU Biodiversity strategy (EC, 2013; EC, 2019b; EC, 2020).
Section 2.2.8. Greening urban and peri-urban areas of the EU Biodiversity Strategy
to 2030 makes explicit reference to UGI, stating: ’... This strategy aims to ... stop
the loss of green urban ecosystems. The promotion of healthy ecosystems, green
infrastructure and nature-based solutions should be systematically integrated into
urban planning, including in public spaces, infrastructure, and the design of
buildings and their surroundings’ (EC, 2020, p. 13).

Urban green spaces provide a broad range of benefits through the maintenance of
ecological function and by contributing to the enhancement of biodiversity
(Benedict et al., 2006; Maes et al., 2020). Strategically deployed and managed
UGI can be multi-functional, providing a wide range of regulating and provisioning
ecosystem services alongside a range of cultural and social values. Some of the
ecosystem services provided by green space that are particularly relevant in urban
areas include air quality and microclimate regulation, protection against flooding,
pollination, recreation and other cultural services (Haase et al., 2014).

The quantity, quality and distribution of green-blue areas is particularly important


for urban ecosystems, human well-being and social cohesion (Raymond et al.,
2017; Sinnet, 2017; Tzoulas et al., 2017). The benefits provided by UGI are
strongly related to other challenge areas. The objective of the Green Space
Management indicators identified herein is to provide a means to assess the
quantity, quality and distribution of green space within cities and their availability
for citizens. Quantity and distribution of UGI are measured considering different
typologies of urban green areas and using as a reference value the total surface of
the city or the total population. The quality of UGI is reported using indicators
related to soil, vegetation, water condition, capacity to provide local food.

The availability of UGI for citizens is measured in terms of accessibility and can be
combined with other indicators to understand users’ preferences and behaviours,
and the availability of facilities that support nature-based activities. Numerous
methods are available to evaluate green space accessibility (Handy and Niemeier,
1997; Páez et al., 2012). Herein, we propose two approaches:

• A relatively simple method that can be easily applied at district and


municipal level and implements parameters recommended by the World
Health Organisation (WHO, 2016; WHO, 2017); and,

• A more complex potential accessibility measure which considers the


cumulative opportunities for nature based recreation and the probability
to reach them according to a function of the distance (Páez et al., 2012).

Other important indicators of Green Space Management, shown herein under


Additional indicators, provide an overview of urban land use intensity considering,

138
for example, land use types and changes, surface sealing (Maes et al., 2019) and
local networks of pedestrian and bicycle paths.

Table 4-4. Indicators related to Green Space Management classified as structural (S), process focused (P)
or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Green space
7.1
accessibility
% O ● ●

Share of green
7.2
urban areas
Number (0-1) O ● ●

Soil organic matter


7.3
content
% O ● ● ●

Soil organic matter


7.3.1
index
Number (0-1) O ● ● ●

ADDITIONAL

Ecosystem
8.1
services provision
N/A; descriptive O ● ● ●

Annual trend in
vegetation cover in
8.2
urban green
% O ●
infrastructure

8.3 Edge density m/ha O ● ●

Public green space


8.4
distribution
ha per capita O ● ●

Distribution of blue
8.5
space
% O ● ●

Effective green
infrastructure at
8.6
the urban-rural
% S ●
interface

Hot spot in peri-


8.7 urban green % S ● ●
infrastructure

139
Biotope Area
8.8
Factor
% O ● ● ●

Total vegetation
8.9
cover
% O ● ● ●

Woody vegetation
8.9.1
cover
% O ● ● ●

Non-woody
8.9.2
vegetation cover
% O ● ● ●

8.9.3 Total leaf area m2 O ● ● ●

Diversity of green
8.10
space
unitless O ● ● ●

Stages of forest
stand development
8.11
-Number of class
No. of individuals O ● ● ●
diameter

8.12 Tree regeneration number O ● ● ●

dychotomic
8.13 Canopy gaps
(Yes/No)
O ● ● ●

Tree biomass stock


8.14
change
t/ha/y O ● ● ●

Measured soil
8.15.1
carbon content
t/ha/y O ● ● ●

Modelled carbon
8.15.2
content
t/ha O ● ● ●

Soil carbon to
8.15.3
nitrogen ratio
unitless O ● ● ●

Soil carbon
8.15.4
decomposition rate
% O ● ● ●

Soil matric
8.16
potential
kPa O ● ● ●

8.17 Soil temperature °C O ● ● ●

Soil water holding


8.18
capacity
mm/cm depth O ● ● ●

Plant-available
8.19.1
water
mm/cm depth O ● ● ●

140
Soil Available
8.19.2 Water (SAW) for mm/cm depth O ● ● ●
plant uptake

Vegetation wilting
8.20
point
% O ●

Degree of soil
8.21
saturation
% O ● ● ●

Stemflow
8.22
funnelling ratio
unitless O ● ● ●

8.23 Soil erodibility mm3/ha O ● ● ●

Total predicted soil


8.24
loss
t/ha/y O ● ● ●

Soil
8.25 ecotoxicological Number (0-1) O ● ● ●
factor

8.26 Soil structure unitless S ●

Soil chemical
8.27 fertility/ cation meq/100 g O ●
exchange capacity

8.28 Flammability Index unitless O ●

Community garden
8.29
area
m2 per capita O ● ●

Food production in
8.30 urban allotments t/ha/y O ● ●
and NBS

Recreational
opportunities
8.31
provided by green
Interactions/week O ● ● ●
infrastructure

ESTIMAP nature-
8.31.1
based recreation
% O ● ● ●

Number of visitors
8.31.2
8.31.3
to recreational No. O ● ● ●
areas

Purpose of visits to
8.31.3
recreational areas
unitless O ● ● ●

Frequency of use
8.31.4 of green and blue h/week O ● ● ●
spaces

141
Activities allowed
8.31.5 in recreational No. S ●
areas

Visual access to
8.32
green space
Number (0-4) O ● ●

Time spent viewing


8.32 green space from Number (0-3) O ● ●
residence each day

8.32.1 Viewshed km2 O ● ●

Satisfaction with
8.32 green and blue Number (1-5) O ● ● ●
spaces

Betweenness
8.34
centrality
unitless O ● ●

Proportion of road
network dedicated
8.35
to pedestrians
% S ●
and/or bicyclists

New pedestrian,
8.35.1 cycling and horse km O ● ●
paths

Sustainable
8.35.2 transportation Number S ●
modes allowed

Links between
8.36 urban centres and Number S ●
NBS

8.37 Walkability Number O ● ● ●

% use class A, N,
8.38 Land composition
D, M
O ● ●

Land use change


8.39 and green space various O ● ●
configuration

8.40 Soil sealing % O ● ●

Ambient pollen
8.41
concentration
Number O ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

142
4.2.5 Biodiversity Enhancement

The fragmentation of green space is a significant impact of urbanisation and can reduce
intra- and inter-species connectivity, leading to a loss of biodiversity. Thus, the structural
and functional connectivity of natural areas (green and blue spaces) are key among
Recommended indicators of biodiversity (indicators 9.1.1 and 9.1.2). Several indicators
are recommended related to the presence of native non-native or alien invasive species
(e.g., 9.2, 9.3 and 9.3.1). These indicators strongly support biodiversity initiatives
focused on the re-introduction or maintenance of local fauna and flora.

Both the Shannon Diversity Index (9.4) and Shannon Evenness Index (9.5) are
recommended indicators of biodiversity. The Shannon Diversity Index is commonly used
to evaluate species diversity within a defined area. Whilst the Shannon Diversity Index
does not qualify whether the species present are native, non-native or alien invasive, it
accounts for the number of different species observed within a given space and their
relative abundances. The Shannon Evenness Index provides information about the
relative number of individuals of each species in a given area.

Numerous additional indicators of biodiversity can support evaluation of the complexity


and multidimensionality of local ecosystems in order to underpin spatial planning,
prioritise sites for interventions and assess the impacts of NBS initiatives on existing
green networks.

Table 4-5. Indicators related to Biodiversity Enhancement classified as structural (S), process focused (P)
or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Structural
9.1 connectivity of
9.1.1 urban green and
various O ● ●
blue spaces

Functional
9.1 connectivity of
9.1.2 urban green and
various O ● ● ●
blue spaces

Number of native
9.2
species
Number O ● ● ●

Number of non-
9.3 native species Number O ● ● ●
introduced

Number of invasive
9.3.1
alien species
Number O ● ● ●

143
Species diversity
9.4 within a defined Number O ● ● ●
area

Number of species
9.5 within a defined Number O ● ● ●
area

ADDITIONAL

Proportion of natural
10.1 areas within a defined % O ● ●
urban zone

Area of habitats
10.2
restored
ha O ● ● ●

Shannon Diversity Number


10.3
Index of habitats (unitless)
O ● ● ●

Abundance of
10.3.1 ecotones/ Shannon unitless O ● ● ●
diversity

10.4 Length of ecotones km O ● ● ●

Publicly accessible
10.5 green space % O ● ● ●
connectivity

10.6 Ecological integrity % O ● ● ●

Proportion of
10.7
protected areas
% O ●

Sites of community
importance and
10.7.1
special protection
ha O ●
areas

Article 17 habitat
10.7.2
richness
No./grid O ● ● ●

Number of veteran
10.8
trees per unit area
No./ha O ● ● ●

Quantity of dead
10.9
wood per unit area
m3/ha O ● ● ●

Forest habitat
fragmentation –
10.10
effective mesh
1/ha O ● ● ●
density

Extent of habitat for


10.11 native pollinator ha O ● ● ●
species

144
10.12 Polluted soils ha O ● ●

10.13 Food web stability unitless O ● ● ●

Carbon and nitrogen


10.14
cycling in soil
t/ha/y O ● ● ●

10.15 Equivalent used soil m3 O ●

Number of
10.16 conservation No. O ● ● ●
priority species

Article 17 species
10.17
richness
No./grid O ● ● ●

Number of native
10.18 bird species within a No./ha O ● ● ●
defined urban area

Species diversity -
10.19
general
No. O ● ● ●

City Biodiversity
10.19.1
Index
% O ● ● ●

Bird species
10.20
richness
No./grid O ● ● ●

Animal species
10.21
potentially at risk
No./ha O ● ● ●

Typical vegetation
10.22
species cover
% O ● ● ●

Pollinator species No./ha or


10.23
presence %
O ● ● ●

Biodiversity
10.24
conservation
various O ● ● ●

Metagenomic
10.25
mapping
unitless O ● ● ●

Abundance of Number
10.25.1
functional groups (unitless)
O ● ● ●

10.25.2 Diversity of Number


10.25.3 functional groups (unitless)
O ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

145
4.2.6 Air Quality

A number of factors threaten the quality of life in European cities and in most of
the world. The drivers include increasing pollution levels, urban heat islands,
flooding and extreme events related to climate change, as well as decreased
biodiversity (Grimm et al., 2008). These can have detrimental effects for human
health and well-being.

Air quality is a major concern worldwide, particularly in urban areas, due to its
direct consequences on human health, plants, animals, infrastructure and
historical buildings (among others). In the political agenda, air quality issues
can be coupled with climate change mitigation policies, since many actions aimed
at air quality improvement involve a concurrent reduction of greenhouse gas
(GHG) emissions. This is the case, for example, of reductions of fossil fuel
combustion since its derived emissions contain CO2 and other GHGs and
pollutants directly affecting human health. Nevertheless, measures to improve
urban air quality and mitigate climate change tend to be considered separately
even though many pollutants affect both environmental impacts.

The emission of the traditional air quality pollutants (AQPs) either direct or
indirectly as a result of atmospheric chemistry, affect the concentrations of
several climate pollutants. At the same time, the increase of air temperature due
to global warming affects the concentrations of the AQPs. Some AQPs, such as
ozone (O3), are also GHGs. These interactions between them are complex and
can both enhance and mitigate global warming. Accordingly, a large number of
abatement measures are beneficial for mitigating both impacts; however, there
are some measures that may be beneficial for mitigating climate change but
increase emissions of the key urban air pollutants, or vice versa.

Policies to reduce climate change and improve urban air quality have
generally been considered in isolation, with more importance being paid to the
mitigation of climate change than to urban air quality over recent years. In the
long term, large reductions in both AQPs and GHGs are necessary to mitigate
climate change and improve public health. Therefore, priority should be given to
measures where there are clear co-benefits such as energy conservation
measures. However, large emissions reductions from this type of measures can
be difficult to achieve and there will continue to be a need to use legislation to
force the adoption of low AQP emitting technologies despite some CO2 penalties.

Fuel switching to renewable fuels offers a huge potential for co-benefits, with only
biomass and biofuels being problematic in terms of indirect GHG emissions from
land use changes and higher emissions of particulate matter (PM) from solid
biomass and gaseous pollutants from some liquid biofuel blends (Querol et al.,
2016).

Air pollution is a local, pan-European and hemispheric issue. Air pollutants


released in one country may be transported in the atmosphere, contributing to
or resulting in poor air quality elsewhere.

Particulate matter, nitrogen dioxide and ground-level ozone, are now


generally recognised as the three pollutants that most significantly affect human

146
health. Long-term and peak exposures to these pollutants range in severity of
impact, from impairing the respiratory system to premature death. Around 90%
of city dwellers in Europe are exposed to pollutants at higher concentrations than
the air quality levels deemed harmful to health. For example, fine particulate
matter (PM2.5) in air has been estimated to reduce life expectancy in the EU by
more than eight months. European Union legislation sets both short-term
(hourly/daily) and long-term (annual) air quality standards 47 (Directive
2008/50/EU). This is reflected in and addressed by the Recommended indicators
(11.1–11.3).

Air pollution also damages our environment. Problems such as acidification


was substantially reduced between 1990 and 2010 in Europe's sensitive
ecosystem areas that were subjected to acid deposition of excess sulphur and
nitrogen compounds. Less progress was made in environmental problematics
such as eutrophication, which is caused by the input of excessive nutrients into
ecosystems. The area of sensitive ecosystems affected by excessive atmospheric
nitrogen diminished only slightly between 1990 and 2010. High ozone
concentrations also cause crop damage is caused. Most agricultural crops are
exposed to ozone levels that exceed the EU long-term objective intended to
protect vegetation. This notably includes a significant proportion of agricultural
areas, particularly in southern, central and eastern Europe.

The Additional indicators of Air Quality focus more specifically on ambient air
pollutant concentration, and the related aspects, such as pollutant removal by
vegetation and associated health aspects.

Table 4-6. Indicators related to Air Quality classified as structural (S), process focused (P) or outcome-
based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Number of days
during which
ambient air pollution
concentrations in
the proximity of the
NBS (PM2.5, PM10, No. of
11.1
O3, NO2, SO2, CO days
O ● ● ●
and/or PAHs
expressed as
concentration of
benzo[a]pyrene)
exceeded threshold

47
http://ec.europa.eu/environment/air/quality/standards.htm;
http://ec.europa.eu/environment/basics/health-wellbeing/noise/index_en.htm

147
values during the
preceding 12
months

Proportion of
population exposed
to ambient air
pollution (PM2.5,
PM10, O3, NO2, SO2,
CO and/or PAHs
11.2 expressed as % O ● ● ●
concentration of
benzo[a]pyrene) in
excess of threshold
values during the
preceding 12
months

Good,
Fair,
Moderate,
European Air Quality
11.3
Index
Poor, Very O ● ● ●
Poor,
Extremely
Poor

ADDITIONAL

Removal of
atmospheric
12.1 pollutants by kg/ha/y O ● ● ●
vegetation (leaves,
stems and roots)

Total particulate
12.2 matter removed by kg/ha/y O ● ● ●
NBS vegetation

Modelled O3, SO2,


NO2 and CO capture/
12.3
removal by
kg/ha/y O ● ● ●
vegetation

12.3.1 Total leaf area m2 O ● ● ●

NOx and PM in PM- µg/m3


12.4
gaseous releases NOx - ppb
O ●

Ambient pollen
12.5
concentration
Number O ● ● ●

Trends in emissions
12.6
of NOx and SOx
µg/m3 O ● ● ●

Concentration of
particulate matter
12.7 (PM10 and PM2.5), µg/m3 O ● ● ●
NO2, and O3 in
ambient air

148
Concentration of
particulate matter
(PM2.5 and PM10) at
12.8
respiration height
µg/m3 O ● ● ●
along roadways and
streets

Mean level of
12.9 exposure to ambient µg/m3 O ● ● ●
air pollution

Morbidity due to
12.10
poor air quality
No./y O ● ● ●

Mortality due to poor


12.10
air quality
No./y O ● ● ●

Years of Life Lost


12.10 due to poor air y O ● ● ●
quality

Avoided costs for air


12.11 pollution control € O ● ● ●
measures


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.7 Place Regeneration

Urban expansion and growth bring countless opportunities and challenges for
cities, rendering place regeneration a significant priority while bringing the
notions of environmental quality and sustainable development to the forefront.
Urban regeneration is seen as a response to the forces pressuring cities to adapt
by addressing decline and increasing the resources for sustainable growth. Urban
regeneration reflects a comprehensive and integrated vision and action which
leads to the resolution of urban problems and which seeks to bring about a lasting
improvement in the economic, physical, social and environmental condition of an
area that has been subject to change (Roberts and Sykes, 2000).

In line with the state-of-the-art in the field of sustainable place regeneration, all
indicators listed here – both recommended and Additional - should be analysed
and applied with consideration for the specific context that defines regeneration
actions at city level, at any given time, the history of a city or area, previous
nature-based initiatives and their impact, as well as other particular issues and
opportunities presented by a town or city.

149
Table 4-7. Indicators related to Place Regeneration classified as structural (S), process focused (P) or
outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Derelict land
13.1
reclaimed for NBS
ha O ●

Quantity of blue-
13.2 green space (as a Number (0-1) O ● ●
ratio to built form)

Perceived quality of
urban blue-green
spaces
(accessibility,
13.3
amenities, natural
various O ● ●
features, incivilities
and recreational
facilities)

Place attachment:
13.4 Place identity or O ● ● ●
“sense of place”

Recreational value
13.5 of public green various O ● ● ●
space

NBS incorporated in
building design /
13.6 incorporation of Number (0-5) P ●
environmental
design in buildings

Cultural heritage
13.7
protection
Number (0-5) P ●

ADDITIONAL

Share of green
14.1
urban areas
% O ● ●

% use class
14.2 Land composition
A, N, D, M
O ● ● ●

14.3 Land take index % O ●

Area devoted to
14.4
roads
Number (0-1) O ● ●

150
Traditional
14.5 knowledge and Yes/No O ● ● ●
uses reclamation

Traditional events
14.6 organised in NBS No. O ● ●
areas

Social active
14.7
associations
No. S ● ● ●

Direct economic
activity: Retail and
14.8 commercial activity % O ● ●
in proximity to
green space

Direct economic
activity: Number of
No. of
new businesses
14.9
created and gross
businesses O ● ●
and €
value added to local
economy

Social return on
14.10
investment
€/€ O ●

14.11 Population mobility % O ● ● ●

14.12 Population growth % O ● ● ●

Proportion of
14.13
elderly residents
% O ● ● ●

14.14 Areal sprawl m2/m2 O ●

Access to public
14.15
amenities
various O ● ●

Average distance of
natural resources
14.16 from urban centres/ km O ● ●
train station/ public
transport

Natural and cultural


14.17
site availability
km2 O ● ●

Historical and
14.18
cultural meaning
unitless O ● ● ●

Cultural value of
14.19
blue-green spaces
various O ● ●

Opportunities for
14.20
tourism
No./year O ● ●

151
Building structure – Dimensionless
14.21
Urban form (0-140)
P ●

Material used
14.22
coherence
Yes/No P ●

Techniques used
14.23
coherence
Yes/No P ●

Design for sense of


14.24
place
Number (0-5) P ● ●

14.25 Viewshed km2 O ● ●

Scenic routes and


14.26
landmarks created
No. O ● ●

Scenic paths
14.27
created
km O ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.8 Knowledge and Social Capacity Building for Sustainable Urban


Transformation

Environmental education opportunities are envisioned as a significant indicator of


urban resources for associational involvement in nature-based solutions, and of
communal contexts for building trust. Although not all environmental education
programs have the potential to generate social capital among participants (e.g.,
classroom instruction), there are forms that can foster social connectivity, trust,
and associational and volunteer involvement. Examples of such programs include
those that incorporate collective opportunities for volunteer and associational
involvement around stewardship, like community gardening and tree planting, or
those that incorporate opportunities for intergenerational learning and collective
decision-making, like place-based learning, school-community partnership for
sustainability, environmental action, action competence, community-based
natural resource management, social-ecological systems resilience) (Krasny et
al., 2015).

The Recommended indicators listed here have been extensively researched as


significant dimensions playing a role in green and pro-environmental behaviour,
NBS impact, and foreseeable sustainability (Derr, 2017; Hedefalk et al., 2015;
Kudryavtsev et al., 2012; Varela-Candamio et al., 2018). The Additional
indicators provide further the means and methods to explore various dimensions
of sustainable urban societal transformation.

152
Table 4-8. Indicators related to Knowledge and Social Capacity Building for Sustainable Urban
Transformation classified as structural (S), process focused (P) or outcome-based (O) indicators and their
general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Citizen involvement
15.1 in environmental No. of people O ● ● ●
education activities

Social learning
Qualitative
regarding
15.2
ecosystems and
data O ● ● ●
(dimensionless)
their functions

Pro-environmental
15.3
identity
O ● ● ●

Pro-environmental Number (0-


15.4
behaviour 168)
O ● ● ●

ADDITIONAL

Children involved in
16.1 educational No./y O ● ● ●
activities

Engagement with Qualitative


16.2 NBS sites and data P ● ● ●
projects (dimensionless)

Number
16.3 Mindfulness
(0-3)
O ● ● ●

Proportion of
schoolchildren
16.4
involved in
% O ●
gardening

Citizens’ awareness
regarding urban Number
16.5
nature and (0-5)
O ● ● ●
ecosystem services

No. activities;
Green intelligence No. attendees;
16.6
awareness No.
O ● ● ●
publications

Positive
16.7
environmental
S, O ● ● ●

153
attitudes motivated
by contact with
NBS

Urban farming
Qualitative
educational and/or
16.8
participatory
data O ● ●
(dimensionless)
activities


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.9 Participatory Planning and Governance

The implementation and scaling of nature-based solutions requires new forms of


planning and governance approaches. In particular, nature-based solutions’
planning and governance need to embrace experimental approaches for
innovation and continuous learning, institutional space for cross-sectoral dialogue
and collaboration and citizen participation (Davies and Lafortezza, 2019;
Frantzeskaki et al., 2019; Kabisch et al., 2017). Citizen participation in
environmental decision-making is extremely valuable, underscoring the
importance of careful consideration of dynamic participation processes through
all the stages of an urban greening project in order to harness the individual and
collective empowering potential of participatory practices (Feldman and
Westphal, 2000). Participatory planning and governance are advocated to
enhance social, political and financial support of the nature-based solution (EC,
2016; Frantzeskaki and Kabisch, 2016; Pauleit et al., 2017).

The recommended indicators capture these cardinal dimensions and processes,


paving the way for a dynamic assessment framework that accounts for processual
variables (e.g., empowerment, trust in decision-making) as well as changes in
existing planning and governance approaches (e.g., new partnerships and policy
learning) (see also Calliari et al., 2019). The additional indicators further explore
relevant participatory processes by examining citizen/stakeholder participation in
NBS planning and implementation, additionally considering the involvement of
under-represented groups. Further dimensions of innovative governance and
financing actions can be explored alongside the adoption of the climate resilience
strategies that highlight the importance of integrated approaches and stakeholder
involvement.

154
Table 4-9. Indicators related to Participatory Planning and Governance classified as structural (S), process
focused (P) or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Openness of
Number
17.1 participatory
(1-5)
P ● ● ●
processes

Proportion of
citizens involved in
17.1.1
participatory
% P ● ● ●
processes

Sense of
empowerment:
17.2 perceived control O ● ● ●
and influence over
decision-making

Adoption of new
forms of
17.3 participatory No. O ● ● ●
governance: PPPs
activated

Policy learning for


mainstreaming NBS:
17.4
Number of new
No. S ● ● ●
policies instituted

Trust in decision-
Number
17.5 making procedure
(1-5)
O ● ● ●
and decision-makers

ADDITIONAL

Community
Number
18.1 involvement in
(0-5)
P ● ● ●
planning

Citizen involvement
18.1.1 in co-creation/ co- No. P ●
design of NBS

Stakeholder
involvement in co-
18.1.2
creation/ co-design
No. P ●
of NBS

Community
Number
18.2 involvement in
(0-5)
P ● ●
implementation

155
Involvement of
citizens from Number
18.3
traditionally under- (0-5)
P ● ● ●
represented groups

Active engagement
18.4 of citizens in % P ● ● ●
decision-making

Consciousness of Number
18.5
citizenship (0-5)
O ● ● ●

Number of
Number
18.6 governance
(0-5)
S ● ● ●
innovations adopted

Adoption of new
Number
18.7 forms of NBS
(0-5)
O ● ● ●
(co-)financing

Development of a
Number
18.8 climate resilience
(0-7)
O ● ● ●
strategy (extent)

Number
Alignment of climate
(0-5)
resilience strategy
18.9
with UNISDR-
across O ● ● ●
117
defined elements
categories

Adaptation of local
plans and Number
18.10
regulations to (0-5)
O ● ● ●
include NBS

Perceived ease of Number


18.11
governance of NBS (0-5)
O ● ● ●

Diversity of
18.12 stakeholders % P ● ● ●
involved

Transparency of co- Number


18.13
production (1-5)
P ● ● ●

Activation of public-
18.14
private collaboration
No. O ● ● ●

Reflexivity:
18.15 identified learning No. P ● ● ●
outcomes

Facilitation skills for Number


18.16
co-production (1-5)
P ● ● ●

Number
18.17 Procedural fairness
(1-5)
P ● ● ●

Number
18.18 Strategic alignment
(1-5)
P ● ● ●

156
Reflexivity: time for
18.19
reflection
No. P ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.10 Social Justice and Social Cohesion

Social cohesion has been long proved to represent an important resource for
long-term environmental sustainability in that socially cohesive communities tend
to be more supportive of environmentally sustainable attitudes and behaviours
compared with those communities where social cohesiveness is weaker (Uzzell et
al., 2002). Bridging social capital’s (indicator 19.1.1) impact on collective
initiatives like nature-based solutions can be far-reaching, as it allows different
groups to share and exchange information, ideas and innovation and builds
consensus among the groups representing otherwise diverse interests.
Conversely, bonding social capital (indicator 19.1.2) fulfils an important social
function by providing the norms and trust that facilitate the kind of collaborative
action required by initiatives like NBS.

Trust, solidarity, tolerance, and respect are generally understood as


manifestations of a cohesive society, one that works towards the well-being of all
the members, that is, towards the common good. While the benefits of
communitarian social capital depend upon basic structural factors (of which
inequality, level of education of the population and its ethnic-racial composition
are considered most important), trust, solidarity, tolerance, and respect
(indicators 19.3-19.5) are cardinal dimensions of the process of creating or
building social capital which enables people to expect good from others
(reciprocity) and to act on behalf of others in order to create a better future for
all (Cloete, 2014).

Moreover, whilst good governance has a significant impact on social cohesion by


increasing trust, tolerance, and acceptance of diversity, creating trust and
guaranteeing reciprocity through concurrent values and abiding to norms that
guide the process of participation in networks are, in fact, acts that fall into the
realm of individual responsibility. It seems that people with values like honesty,
trustworthiness, integrity, who care for their fellow humans, are likely to create
social capital that could lead to the formation of public good (Cloete, 2014).
Therefore, trust, solidarity, tolerance, and respect are considered fundamental
resources in the inception, implementation, and potential success of any
collective initiatives like nature-based solutions.

All things considered, the Recommended indicators included here address the
main dimensions pertinent to state-of-the-art research of nature-based solution
and their role in creating social capital and fostering global priorities oriented

157
towards social cohesion and social justice. The Additional indicators focus on the
supplementary details, including perceived social interactions, safety and
inclusion, and crime.

Table 4-10. Indicators related to Social Justice and Social Cohesion classified as structural (S), process
focused (P) or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Bridging– quality of
interactions within
19.1.1
and between social
O ● ● ●
groups

Bonding – quality of
interactions within
19.1.2
and between social
O ● ● ●
groups

Inclusion of different
Number
19.2 social groups in NBS
(0-5)
P ● ● ●
co-co-co processes

Trust within the


19.3
community
O ● ● ●

Solidarity among
19.4
neighbours
O ● ● ●

Tolerance and
19.5
respect
O ● ● ●

Availability and
19.6 equitable distribution map O ● ● ●
of blue-green space

ADDITIONAL

20.1 Linking social capital O ● ● ●

Number
Perceived social (0-5)
20.2
interaction across 4
O ● ● ●
categories

Quantity and quality


20.3
of social interaction
Frequency O ● ● ●

158
Number
Perception of socially (0-5)
20.4.1
supportive network across 5
O ● ● ●
categories

Perceived social Number


20.4.2
support (0-4)
O ● ● ●

Perceived social Number


20.5
cohesion (0-4)
O ● ● ●

Perceived ownership Number


of space and sense of (0-5)
20.6
belonging to the across 2
O ● ● ●
community categories

Proportion of
Number
20.7 community who
(0-5)
O ● ●
volunteer

Proportion of target
20.8 group reached by an % O ● ● ●
NBS project

Perceived personal Number


20.9
safety (0-5)
O ● ● ●

Perceived safety of
20.10
neighbourhood
O ● ● ●

Number of violent
incidents, nuisances No. per
20.11
and crimes per 100 000
O ● ● ●
100 000 population

20.12 Realised safety O ● ● ●

Area easily accessible


20.13 for people with km2 O ● ● ●
disabilities

Change in property
20.14
incomes
% O ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.11 Health and Wellbeing

The effects of climate change, such as heatwaves, lead to urban areas becoming
increasingly uncomfortable, with vulnerable members of society feeling such impacts

159
the most 48. In the heat wave of summer 2003 in Europe for example, more than
70 000 excess deaths were recorded (Robine et al., 2008).

High temperatures also raise the levels of ozone and other pollutants in the air that
exacerbate cardiovascular and respiratory disease 49. Air quality (see section 4.2.6)
is also a major concern worldwide, particularly in urban areas, due to its direct
consequences on human health, plants, animals, infrastructure and historical
buildings (among others). Increasing evidence supports the idea that ecological
features such as the diurnal cycles of light and day, sunlight exposure, seasons, and
geographic characteristics of the natural environment such as altitude, latitude, and
green spaces are important determinants of cardiovascular health and cardiovascular
disease (CVD) risk (Bhatnagar, 2017). Some of the beneficial cardiovascular effects
of greenery might relate to a decrease in the levels of local air pollution, increased
proximity to walking spaces, or lower levels of mental stress (Bhatnagar, 2017). With
an abundance of convenient, palatable, energy dense foods and increasingly fewer
demands for physical activity in usual lifestyles, the contemporary environment
enables the energy balance to be tipped in favour of weight gain (obesogenic
environment) (Bhrem and D'Alessio, 2014). In adults, obesity is associated with
increasing risk of cardiovascular disease, type 2 diabetes, and all-cause mortality.
Most of the associated mortality and morbidity is mediated through major chronic
diseases related to obesity, such as cardiovascular disease, diabetes, and cancer
(Bhrem and D'Alessio, 2014). Overweight children face a greater risk of a host of
problems, including type 2 diabetes, high blood pressure, high blood lipids, asthma,
sleep apnoea, chronic hypoxemia (too little oxygen in the blood), early maturation,
and orthopaedic problems (Samuels, 2004). They also suffer psychosocial problems,
including low self-esteem, poor body image, and symptoms of depression (Samuels,
2004). This is highlighted by Recommended indicators (21.1, 21.5, 21.6).

Climate change means that floods are also increasing in frequency and intensity, and
the frequency and intensity of extreme precipitation is expected to continue to
increase throughout the current century (IPCC, 2014). A decrease in experienced
nature is one aspect of urbanisation that has drawn researchers’ attention with the
purpose of developing methodologies to explore the affective and cognitive benefits
of nature experience, and demonstrate the psychological benefits of our exposure
to/engagement with nature (Bratman et al., 2015). The mental health benefits of
urban green space have been highlighted by a growing body of knowledge and
empirical evidence attesting to the complex interplay among stress responses,
neighbourhood conditions, and health outcomes (Beyer et al., 2014; Frumkin et al.,
2017; Hartig et al., 2014). More greenery in the neighbourhood was linked to lower
levels of depression, anxiety, and stress (Beyer et al., 2014; Pope et al., 2015).
Moreover, mental restoration and relaxation from leisure activities (e.g., walks in
parks vs. walks in urban settings, gardening) pursued in the nature and green space
have been studied as strong evidence of mental health benefits consequent to nature
experience (Aspinall et al., 2013; Bratman et al., 2015; Braubach et al., 2017;,
Hartig et al., 2014; van der Berg and Custers, 2011). These aspects are addressed
in Recommended indicators 21.2–21.4, and 21.6.

48 Climate change, justice and vulnerability. http://bit.ly/16STKgy


49
http://www.who.int/news-room/fact-sheets/detail/climate-change-and-health

160
Numerous authors emphasize that modern urban wellbeing challenged by chronic
stress (indicator 21.2) and insufficient physical activity can be healthily nurtured by
natural environment exposure, which promotes mental and physical health and
reduces morbidity and mortality in urban residents by providing psychological
relaxation (indicators 21.3, 21.4) and stress alleviation, enhancing immune function,
stimulating social cohesion, supporting physical activity (indicator 21.1), and
reducing exposure to air pollutants, noise and excessive heat (Braubach et al., 2017;
Hartig et al., 2014).

These health and wellbeing benefits are important not just at the individual level, but
if implemented widely they could save expenditure on health care. Increasing the
extent and improving the quality of green spaces in areas of cities where health
outcomes are poor could also play an important role in addressing multiple
deprivations.

Research on complex/multi-dimensional relationship between nature


connectedness/nature affiliation (i.e., affective, cognitive and experiential factors
related to our belonging to the natural world) and wellbeing indicate that exposure
to elements of the natural world affects our well-being by boosting our positive affect,
by eliciting feelings of ecstasy, respect, and wonder, by fostering feelings of comfort
and friendliness, by heightening our intrinsic aspirations and generosity, and by
increasing our vitality (Capaldi et al., 2014; Howell and Passmore, 2013), highlighted
in Recommended indicators 21.3 and 21.4, and Additional indicators 22.11, 22.13,
and 22.15.

The Additional indicators of NBS impacts on Health and Wellbeing focus on evaluating
health and wellbeing aspects in relation to noise, heat and air pollution, and exploring
psychological and chronic stress changes, including anxiety, in greater depth.

Table 4-11. Indicators related to Health and Wellbeing classified as structural (S), process focused (P) or
outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Level of outdoor
21.1
physical activity
O ● ●

Level of chronic
Number
21.2 stress (perceived
(0-4)
O ● ● ●
stress)

General wellbeing Number


21.3
and happiness (0-7)
O ● ● ●

161
Self-reported mental Number
21.4
health and wellbeing (1-6)
O ● ● ●

Prevalence of
21.5 cardiovascular % O ● ●
disease

Incidence of
21.5 cardiovascular % per year O ● ●
disease

Number
21.6 Quality of life
(1-5)
O ● ● ●

ADDITIONAL

Self-reported Minutes
22.1
physical activity per week
O ● ●

% over
three
levels of
Observed physical physical
22.2
activity within NBS activity
O ● ●
(sedentary,
walking, or
vigorous)

Encouraging a Number
22.3
healthy lifestyle (1-5)
O ● ●

Morbidity due to
21.5 cardiovascular No./y O ● ●
disease

Mortality due to
21.5 cardiovascular No./y O ● ●
disease

22.4 Incidence of obesity % per year O ● ●

Heat-related
discomfort:
22.5 Universal Thermal °C O ● ●
Climate Index
(UTCI)

Hospital admissions
due to high No. per
22.6
temperature during 100 000
O ● ●
extreme heat events

No. per
Heat-related
22.7
mortality
1 000 000 O ● ●
per year

Exposure to noise
22.8
pollution
% O ● ●

162
Number
Perceived chronic (1-3)
22.9
loneliness across 3
O ● ● ●
categories

Low,
Moderately
22.10 Somatisation
high, Very
O ● ● ●
high

Number
(0-4)
22.11 Mindfulness
across 12
O ● ● ●
categories

Visual access to Number


22.12
green space (0-4)
O ● ●

Time spent viewing


Number
22.12 green space from
(0-3)
O ● ●
residence each day

Perceived Number
restorativeness of (0-10)
22.13
public green space/ across 4
O ● ● ●
NBS categories

Perceived social Number


22.14
support (0-4)
O ● ● ●

Number
Connectedness to (1-5)
22.15
nature across 14
O ● ● ●
categories

Prevalence of
attention deficit
22.16
hyperactivity
% O ● ●
disorder (ADHD)

Exploratory
22.17 behaviour in O ● ●
children

Mild,
Self-reported
22.18
anxiety
Moderate, O ● ● ●
Severe

Prevalence of
22.19
respiratory diseases
% O ● ● ●

Incidence of
22.19
respiratory diseases
% per year O ● ● ●

Morbidity of
22.19
respiratory diseases
No./y O ● ● ●

Mortality of
22.19
respiratory diseases
No./y O ● ● ●

163
Morbidity due to
22.20
poor air quality
No./y O ● ● ●

Mortality due to
22.20
poor air quality
No./y O ● ● ●

Years of life lost


No. of
22.20 (YoLL) due to poor
years
O ● ● ●
air quality

Prevalence of
22.21 autoimmune % O ● ● ●
diseases

Incidence of
22.21 autoimmune % per year O ● ● ●
diseases

Prevalence of
22.22
chronic stress
% O ● ● ●

Incidence of chronic
22.22
stress
% per year O ● ● ●

Morbidity due to
22.22
chronic stress
No./y O ● ● ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional
interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

4.2.12 New Economic Opportunities and Green Jobs

The economic opportunities that are created by the adoption and implementation of
NBS as a consequence of their social attractiveness and restorative value can be
evaluated using the Recommended indicators 23.2, 23.4–23.6. Indicator 23.2 and
related sub-indicators 23.2.1-23.2.3 provide several different metrics to evaluate
changes in mean land or property value attributable to the implementation of local
NBS. Indicator 23.4 specifically evaluates the use of ground floor building space for
retail, commercial or public purposes in the proximity of NBS, whilst indicator 23.5
examines the gross value added (GVA) to the local economy each year in the area
near implemented NBS. The value of recreational activities occurring in NBS is
addressed by indicator 23.6.

Indicators of new economic opportunities are supported by assessment of the value


of new jobs created per annum (23.3) as a result of new business opportunities and
new jobs in the green sector. Green jobs are those that contribute environmental
benefit. The International Labour Organization (ILO) defines green jobs within three
categories: primary green activities (i.e., organic agriculture, sustainable forestry),
secondary activities (i.e., renewable energy, clean industry, sustainable

164
construction) and tertiary activities (i.e., recycling, sustainable tourism, and
sustainable transport).

There has been a great deal of research on the valuation of the benefits provided by
the natural environment using a wide range of techniques. Indicators supporting the
valuation of urban nature (23.1.1 and 23.1.2) and its ecosystem services enable
quantification of NBS benefits translated into monetary terms. Economic valuation of
NBS benefits provides a much-needed means to inform decision-making.

Additional indicators within the New Economic Opportunities and Green Jobs
challenge area examine indirect economic activity in the area surrounding NBS,
elements of NBS cost-benefit analysis (including the value of hydro-meteorological
risk reduction), social return on investment, the value of NBS-based tourism, and
the impact of local innovation, among others. The indicators identified for the New
Economic Opportunities and Green Jobs challenge area address a relatively broad
range of actions and potential or realised economic consequences.

Table 4-12. Indicators related to New Economic Opportunities and Green Jobs classified as structural (S),
process focused (P) or outcome-based (O) indicators and their general applicability to different types of NBS

Indicator Units Class Applicability to NBS†


No.

Type 1 Type 2 Type 3

RECOMMENDED

Valuation of NBS:
Value of NBS
23.1.1
calculated using
€ O ● ● ●
GI-Val

Economic value of
23.1.2
urban nature
€ O ● ● ●

Mean land and/ or


property value in
23.2
proximity to green
€ O ● ●
space

Change in mean
23.2.1 house prices/ rental € O ● ●
markets

Average land
23.2.2 productivity and €/ha O ● ● ●
profitability

Property
betterment and
23.2.3
visual amenity
€/m2 O ● ●
enhancement

165
Direct economic
23.3 activity: Number of €/year O ● ● ●
new jobs created

Direct economic
activity: Retail and
23.4 commercial activity % O ● ●
in proximity to
green space

Direct economic
activity: Gross
value added to
23.5
local economy from
%/year O ● ● ●
new business
creation

Recreational
23.6
monetary value
€/year O ● ●

Overall economic, Human


23.7 social and health Development O ● ● ●
well-being Index

ADDITIONAL

Indirect economic
activity: number of
24.1 new businesses No./year O ● ●
established in
proximity to NBS

Indirect economic
activity: Value of
24.2 rates paid by €/year O ● ●
businesses in
proximity to NBS

Indirect economic
activity: New Mean
24.3 customers to No./day per O ● ●
businesses in quarter
proximity to NBS

Indirect economic
activity: local
24.4
economy GDP in
€/year O ● ●
proximity to NBS

NBS cost/benefit
24.5
analysis: Initial costs
€ O ● ● ●

NBS cost/benefit
24.6 analysis: €/year O ● ● ●
Maintenance costs

NBS cost/benefit
24.7 analysis: € O ● ● ●
Replacement costs

166
NBS cost/benefit
24.8 analysis: Avoided € O ● ● ●
costs

NBS cost/benefit
24.9 analysis: Payback year O ● ● ●
period

Reduced/ avoided
damage costs from
24.10 hydro €/year O ● ● ●
meteorological risk
reduction

Social return on
24.11
investment (SROI)
€/€ O ● ● ●

Income generated
via application of
green
24.12
administrative
€/year O ● ● ●
policies within
Living Lab district

Subsidies applied
24.13 for private NBS €/year O ● ● ●
measures

Private finance
attracted to the
24.14 NBS site/ private €/year O ● ● ●
investment in the
bioeconomy

Mean no.
24.15 Increase in tourism visitors/day O ● ●
per year

New activities in Number


24.16
the tourism sector (1-5)
O ● ●

Gross profit from €/year per


24.17
nature-based tourism km2
O ● ●

Number of new
24.18
jobs in green sector
% O ● ● ●

Number of new
jobs related to NBS Number
24.19
construction and (1-5)
O ● ● ●
maintenance

New employment in Number


24.20
the tourism sector (1-5)
O ● ●

Turnover in the
24.21
green sector
% O ● ● ●

Employment in
24.22
agriculture
No./ha O ● ● ●

167
Rural Productivity
24.23
Index
€/ha O ● ● ●

Economic value of
the productive
24.24
activities vulnerable
€/km2 O ● ● ●
to risks

No.
24.25 Innovation impact
innovations
O ● ● ●

€/year per
24.26 Income per capita
person
O ● ● ●

24.26 Disposable income €/year per


24.26.1 per capita person
O ● ● ●

Increase in
Upskilling and employment
24.27 related earnings earnings per O ● ● ●
increase person per
year

% in 1 y
24.28 Population mobility % in 2 y O ● ●
% in 5 y

Avoided cost of
24.29
run-off treatment
€/y O ● ● ●

Correction cost of
24.30
groundwater quality
€/m3 O ● ● ●

Dissuasive cost of
24.31
water abstraction
€/m3 O ● ● ●

Average water
24.32
productivity
€/m3 O ● ● ●

New areas made


available for
24.33
traditional
km2 O ● ● ●
productive uses

Value of food
24.34
produced in NBS
€/y O ● ●

Renewable energy
24.35
produced in NBS
kWh/y O ●


Type 1 NBS – minimal or no intervention in ecosystems, with objectives related to maintaining or improving
delivery of ecosystem services within and beyond the protected ecosystems
Type 2 NBS – extensive or intensive management approaches seeking to develop sustainable, multifunctional
ecosystems and landscapes in order to improve delivery of ecosystem services relative to conventional interventions
Type 3 NBS – characterised by highly intensive ecosystem management or creation of new ecosystems

168
4.3 Conclusions

4.3.1 Summary of the indicator framework presented

The Recommended indicators, taken together, are designed to provide a holistic


assessment of the multiple potential co-benefits of NBS. Practitioners are
encouraged to adopt as many of these Recommended indicators as practicable.
Depending upon the specific context, some Recommended indicators may not be
entirely applicable or may require adaptation to the local conditions or to
overcome resource (personnel, equipment, finance) limitations. In such cases,
the Additional indicators presented herein may serve as support, providing
opportunity for monitoring and evaluation framework adaptation and tailoring to
local conditions as necessary.

Critical thinking is required to select the indicators that suit the purpose and the
scope of the NBS assessment strategy. Detailed information regarding the
applicability and requirements for each indicator analysis are presented in the
Appendix.

4.3.2 Emerging concerns and further development needs

There were a number of indicators initially discussed by the members of the


H2020 NBS projects involved in producing this handbook that were ultimately not
included herein due to a lack of consensus regarding assessment methodology.
In many cases, further work is required to validate evaluation methods for a
variety of the NBS forms and functions in order to establish a standardised
procedure for assessment of NBS impact. Outcomes of on-going and future NBS
projects are expected to deliver novel indicators of NBS impact across all societal
challenge areas identified here.

Greater confidence in techniques for evaluation are needed, particularly for


carbon flux measurements from natural ecosystems and heterogeneous urban
areas. Reduction in price of monitoring equipment with technological advances
should make monitoring more accessible and applicable.

Concerning the water management challenge, one of the main concerns is the
identification and development of synergic strategies to safeguard and properly
support ecosystem services. The effective detection of spatial and temporal scales
allows assessing and fostering the ecosystem resilience and sustainability.
Attention should be paid to investigating alternations to flow regime to account
for the uncertainty and non-stationarity of the hydrologic methodologies.
Technological advancement will make monitoring more accessible and applicable,
particularly in relation to automated sampling and analysis, and in-pipe
measurements of low flowrates. Advances in the accessibility of high-resolution
imagery will yield more monitoring options.

For biodiversity assessment, greater standardisation of approaches is needed,


this may come through increased requirement for reporting through legislative

169
and planning processes. There is also a need for indicators that capture the
complexity and diversity of biodiversity evaluation beyond the usual suspects.

Additionally, a wide variety of indicators and methodologies are presented in this


manual, not all of which have been validated to assess large-scale NBS
interventions. In this sense, the results obtained in the current H2020 projects
will serve to guide future projects and implementations in the selection of the
most appropriate in each case. Likewise, it is necessary to consider the impacts
of the COVID-19 pandemic on some of the assessment methodologies presented
in this handbook and Appendix of Methods as some KPIs may require
modifications to the way they are evaluated (e.g., changes to how use of green
spaces is assessed due to local restrictions on movement). In some cases, the
units of the KPIs may be modified to better apply to a specific case study or to
improve the understanding of results.

4.4 References

Aspinall, P., Mavros, P., Coyne, R., and Roe, J., ‘The urban brain: Analyzing outdoor physical activity with
mobile EEG’, British Journal of Sports Medicine, Vol. 49, No 4, 2013, pp. 272-276.

Benedict, M.A. and McMahon, E.T., ‘Green infrastructure’, Island, Washington, D.C., 2006.

Beyer, K.M., Kaltenbach, A., Szabo, A., Bogar, S., Nieto, F.J., and Malecki, K.M., ‘Exposure to neighborhood
green space and mental health: Evidence from the Survey of the Health of Wisconsin’, International
Journal of Environmental Research and Public Health, Vol. 11, No 3, 2014, pp. 3453-3472.

Bhatnagar A., ‘Environmental Determinants of Cardiovascular Disease’, Circulation Research, Vol. 121, No 2,
2017, pp. 162–180.

Brehm, B.J. and D’Alessio, D.A., ‘Environmental factors influencing obesity’, Endotext, MDText.com, Inc.,
South Dartmouth, 2014, pp. 2000. Retrieved from https://www.ncbi.nlm.nih.gov/books/NBK278977/

Bratman, G.N., Hamilton, J.P., Hahn, K.S., Daily, G.C., and Gross, J.J., ‘Nature experience reduces rumination
and subgenual prefrontal cortex activation’, Proceedings of the National Academy of Sciences of the
United States of America, Vol. 112, No 28, 2015, pp. 8567–8572.

Braubach, M., Egorov, A., Mudu, P., Wolf, T., Ward Thompson, C., and Martuzzi, M., ‘Effects of Urban Green
Space on Environmental Health, Equity and Resilience’, Nature-Based Solutions to Climate Change
Adaptation in Urban Areas: Linkages between Science, Policy and Practice, SpringerOpen, Cham,
2017, pp. 187-205.

Calliari, E., Staccione, A., and Mysiak, J., ‘An assessment framework for climate-proof nature-based solutions’,
Science of the Total Environment, Vol. 656, 2019, pp. 691-700.

Capaldi, C.A., Dopko, R.L., and Zelenski, J.M., ‘The relationship between nature connectedness and happiness:
a meta-analysis’, Frontiers in Psychology, Vol. 5, 2014, pp. 976.

Cloete, A., ‘Social cohesion and social capital: Possible implications for the common good’, Verbum et Ecclesia,
Vol. 35, No 3, 2014, Art. no 1331, 6 pp.

Cohen-Shacham, E., Walters, G., Janzen, C., and Maginnis, S. (Eds.), Nature-based Solutions to address global
societal challenges, International Union for the Conservation of Nature, Gland, Switzerland, 2016.

Davies, C. and Lafortezza, R., ‘Transitional path to the adoption of nature-based solutions’, Land Use Policy,
Vol. 80, 2019, pp. 406–409.

Demuzere, M., Orru, K., Heidrich, O., Olazabal, E., Geneletti, D., Orru, H., and Faehnle, M., ‘Mitigating and
adapting to climate change: Multi-functional and multi-scale assessment of green urban
infrastructure’, Journal of Environmental Management, Vol. 146, 2014, pp. 107-115.

Derr, V., ‘Urban green spaces as participatory learning laboratories’, Proceedings of the Institution of Civil
Engineers-Urban Design and Planning, Vol. 171, No 1, 2017, pp. 25-33.

170
Donabedian, A., ‘Evaluating the quality of medical care’, The Millbank Memorial Fund Quarterly, Vol. 44, 1966,
pp. 166-203.

Eggermont, H., Balian, E., Azevedo, J.M.N., Beumer, V., Brodin, T., Claudet, J., Fady, B., Grube, M., Keune,
H., Lamarque, P., Reuter, K., Smith, M., van Ham, C., Weisser, W.W., and Le Roux, X., ‘Nature-based
solutions: New influence for environmental management and research in Europe’, GAIA, Vol. 24, No
4, 2015, pp. 243-248.

Ernst, L., de Graaf-Van Dinther, R.E., Peek, G.J., and Loorbach, D.A., ‘Sustainable urban transformation and
sustainability transitions; conceptual framework and case study’, Journal of Cleaner Production, Vol.
112, 2016, pp. 2988-2999.

European Commission, Green Infrastructure (GI) — Enhancing Europe’s Natural Capital, COM(2013) 249 final,
2013.

European Commission, Towards an EU Research and Innovation policy agenda for Nature-Based Solutions &
Re-Naturing Cities. Final Report of the Horizon 2020 Expert Group on 'Nature-Based Solutions and Re-
Naturing Cities', Publications Office of the European Union, Luxembourg, 2015.

European Commission, Commission Staff Working Document: Guidance on a strategic framework for further
supporting the deployment of EU-level green and blue infrastructure, SWD(2019) 193 final, 2019a.

European Commission, Report from the Commission to the European Parliament, the Council, the European
Economic and Social Committee and the Committee of the Regions: Review of progress on
implementation of the EU green infrastructure strategy, COM(2019) 236 final, 2019b.

European Commission, EU Biodiversity Strategy for 2030: Bringing Nature Back into Our Lives, COM(2020)
380 Final, 2020.

Feldman, R.M. and Westphal, L.M., ‘An agenda for community design and planning: Participation and
empowerment practice’, Sustaining human settlement: a challenge for the new millennium, Urban
International Press, North Shields, Great Britain, 1999, pp. 105-139.

Frantzeskaki, N., ‘Seven lessons for planning nature-based solutions in cities’, Environmental Science and
Policy, Vol. 93, 2019, pp. 101-111.

Frantzeskaki, N. and Kabisch, N., ‘Designing a knowledge co-production operating space for urban
environmental governance—Lessons from Rotterdam, Netherlands and Berlin, Germany’,
Environmental Science and Policy, Vol. 62, 2016, pp. 90-98.

Frumkin, H., Bratman, G.N., Breslow, S.J., Cochran, B., Kahn, P.H., Jr., Lawler, J.J., Levin, P.S., Tandon, P.S.,
Varanasi, U., Wolf, K.L., and Wood, S.A., ‘Nature Contact and Human Health: A Research Agenda’,
Environmental Health Perspectives, Vol. 125, No 7, 2017, pp. 075001.

Gentin, S., Pitkänen, K., Chondromatidou, A.M., Præstholm, S., Dolling, A., and Palsdottir, A.M., ‘Nature-
based integration of immigrants in Europe: A review’, Urban Forestry and Urban Greening, Vol. 43,
2019, pp. 1-8.

Grimm, N.B., Faeth, S.H., Golubiewski, E.N., Redman, C.L., Wu, J., Bai, X., and Briggs, J.M., ‘Global change
and the ecology of cities’, Science, Vol. 319, 2008, pp. 756–760.

Haase, D., Larondelle, N., Andersson, E., Artmann, M., Borgström, S., Breuste, J., Gomez-Baggethun, E.,
Gren, Å., Hamstead, Z., Hanson, R., Kabisch, N., Kremer, P., Langemeyer, J., Lorance Rall, E.,
McPhearson, T., Pauleit, S., Qureshi, S., Schwartz, N., Voigt, A., Wurster, S., and Elmqvist, T., ‘A
Quantitative Review of Urban Ecosystem Service Assessments: Concepts, Models, and
Implementation’, AMBIO, Vol. 43, No 4, 2014, pp. 413–433.

Handy, S.L. and Niemeier, D.A., ‘Measuring accessibility: an exploration of issues and alternatives’,
Environment and Planning A, Vol. 29, No 7, 1997, pp. 1175-1194.

Hartig, T., Mitchell, R., de Vries, S., and Frumkin, H., ‘Nature and Health’, Annual Review of Public Health,
Vol. 35, 2014, pp. 207-228.

Hedefalk, M., Almqvist, J., and Östman, L., ‘Education for sustainable development in early childhood
education: a review of the research literature’, Environmental Education Research, Vol. 21, No 7,
2015, pp. 975-990.

Howell, A.J. and Passmore, H.-A., ‘The nature of happiness: Nature affiliation and mental well-being’, Mental
well-being: International contributions to the study of positive mental health, Springer, New York,
2013, pp. 231–257.

Ibes, D.C., ‘A multi-dimensional classification and equity analysis of an urban park system: A novel
methodology and case study application’, Landscape and Urban Planning, Vol 137, 2015, pp. 122–
137.

171
IPCC, ‘Summary for Policymakers’, Climate Change 2014: Mitigation of CC. Contribution of Working Group
III to the 5th Assessment Report of the IPCC, New York, Cambridge University Press, 2014.

Jahn, T., Bergmann, M., and Keil, F., ‘Transdisciplinarity: Between mainstreaming and marginalization’,
Ecological Economics, Vol. 79, 2012, pp.1-10.

Kabisch, N., Korn, H., Stadler, J., and Bonn, A. (Eds.), ‘Nature-Based Solutions to Climate Change Adaptation
in Urban Areas’, Theory and Practice of Urban Sustainability Transitions, Springer, Cham, 2017.

Korkmaz, C. and Balaban, O., ‘Sustainability of urban regeneration in Turkey: Assessing the performance of
the North Ankara Urban Regeneration Project’, Habitat International, Vol. 95, 2020, pp. 1-14.

Kudryavtsev, A., Krasny, M.E., and Stedman, R.C., ‘The impact of environmental education on sense of place
among urban youth’, Ecosphere, Vol. 3, No 4, 2012, pp. 1-15.

Kweon, B.S., Sullivan, W.C., and Wiley, A.R., ‘Green common spaces and the social integration of inner-city
older adults’, Environment and Behavior, Vol. 30, No 6, 1998, pp. 832-858.

Maes, J., Zulian, G., Günther, S., Thijssen, M., and Raynal, J., Enhancing Resilience of Urban Ecosystems
through Green Infrastructure. Final Report, EUR 29630 EN, Luxembourg, Publications Office of the
European Union, 2019.

Maes, J., Zulian, G., Thijssen, M., Castell, C., Baró, F., Ferreira, A.M., Melo, J., Garrett, C.P., David, N., Alzetta,
C., Geneletti, D., Cortinovis, C., Zwierzchowska, I., Louro Alves, F., Souto Cruz, C., Blasi, C., Alós
Ortí, M.M., Attorre, F., Azzella, M.M., Capotorti, G., Copiz, R., Fusaro, L., Manes, F., Marando, F.,
Marchetti, M., Mollo, B., Salvatori, E., Zavattero, L., Zingari, P.C., Giarratano, M.C., Bianchi, E., Duprè,
E., Barton, D., Stange, E., Perez-Sob, M., van Eupen, M., Verweij, P., de Vries, A., Kruse, H., Polce,
C., Cugny-Seguin, M., Erhard, M., Nicolau, R., Fonseca, A., Fritz, M., and Teller, A., Mapping and
Assessment of Ecosystems and their Services, Urban Ecosystems, Publications Office of the European
Union, Luxembourg, 2016.

Maes, J., Teller, A., Erhard, M., Conde, S., Vallecillo Rodriguez, S., Barredo Cano, J.I., Paracchini, M., Abdul
Malak, D., Trombetti, M., Vigiak, O., Zulian, G., Addamo, A., Grizzetti, B., Somma, F., Hagyo, A.,
Vogt, P., Polce, C., Jones, A., Marin, A., Ivits, E., Mauri, A., Rega, C., Czucz, B., Ceccherini, G., Pisoni,
E., Ceglar, A., De Palma, P., Cerrani, I., Meroni, M., Caudullo, G., Lugato, E., Vogt, J., Spinoni, J.,
Cammalleri, C., Bastrup-Birk, A., San-Miguel-Ayanz, J., San Román, S., Kristensen, P., Christiansen,
T., Zal, N., De Roo, A., De Jesus Cardoso, A., Pistocchi, A., Del Barrio Alvarellos, I., Tsiamis, K.,
Gervasini, E., Deriu, I., La Notte, A., Abad Viñas, R., Vizzarri, M., Camia, A., Robert, N., Kakoulaki,
G., Garcia Bendito, E., Panagos, P., Ballabio, C., Scarpa, S., Montanarella, L., Orgiazzi, A., Fernandez
Ugalde, O., and Santos-Martín, F., Mapping and Assessment of Ecosystems and their Services: An EU
ecosystem assessment, EUR 30161 EN, Publications Office of the European Union, Luxembourg, 2020.

McCormick, K., Anderberg, S., Coenen, L., and Neij, L., ‘Advancing sustainable urban transformation’, Journal
of Cleaner Production, Vol. 50, 2013, pp. 1-11.

Páez, A., Scott, D.M., and Morency, C., ‘Measuring accessibility: positive and normative implementations of
various accessibility indicators’, Journal of Transport Geography, Vol. 25, 2012, pp. 141–153.

Pauleit, S., Zölch, T., Hansen, R., Randrup, T.B., and van den Bosch, C.K., ‘Nature-based solutions and
climate change–four shades of green’, Nature-Based Solutions to Climate Change Adaptation in
Urban Areas, Springer, Cham, 2017, pp. 29-49.

Pope, D., Tisdall, R., Middleton, J., Verma, A., Van Ameijden, E., Birt, C., and Bruce, N.G., ‘Quality of and
access to green space in relation to psychological distress: results from a population-based cross-
sectional study as part of the EURO-URHIS 2 project’, European Journal of Public Health, Vol. 28,
No 1, 2015, pp. 35-38.

Querol, X., Karanasiou, A., Amato, F., Vasconcelos, C., Alastuey, A., Viana, M., Moreno, T., Plana, F., Perez,
N., Cabañas, M., Bartoli, R., Martinez, S., Sosa, M., Montfort, E., Celades, I., Escrig. A., Sanfelix, V.,
Gomar, S., Harrison, R., Holman, C., Beddows, D., Harding, M., Eleftheriadis, K., Diapouli, L.,
Vratoulis, S., Gini, M., Bairaktari, E., Dalaina, S., Galifianakis, V., Lucarelli, F., Nava, S., Calzolai, G.,
Udisti, R., Becagli, S., Traversi, R., Severi, M., Borselli, S., Giannoni, M., Alves, C., Pio, C., Nunes, T.,
Tarelho, L., Duarte, M., Cerqueira, M., Vicente, E., Custódio, D., Pinto, H., Gianelle, V.L., and Colombi,
C., ‘Air Quality and Climate Change Abatement: Synergies and Conflicts, Report 23, 12/2016’, LIFE
AIRUSE (LIFE11 ENV/ES/584), 2016.

Raymond, C.M., Frantzeskaki, N., Kabisch, N., Berry, P., Breil, M., Nita, M.R., Geneletti, D., and Calfapietra,
C., ‘A framework for assessing and implementing the co-benefits of nature-based solutions in urban
areas’, Environmental Science and Policy, Vol. 77, 2017, pp. 15-24.

Raymond, C.M., Gottwald, S., Kuoppa, J., and Kyttä, M., ‘Integrating multiple elements of environmental
justice into urban blue space planning using public participation geographic information systems’,
Landscape and Urban Planning, Vol. 153, 2016, pp. 198-208.

172
Roberts, P. and Sykes, H. (Eds.), Urban regeneration: a handbook, London, SAGE Publications Ltd, 2000.

Robine, J.-M., Cheung, S.L.K., Le Roy, S., Van Oyen, H., Griffiths, C., Michel, J.-P., and Herrmann, F.R., ‘Death
toll exceeded 70,000 in Europe during the summer of 2003 / Plus de 70 000 décès en Europe au cours
de l’été 2003’, Comptes Rendes Biologies, Vol. 331, No 2, 2008, pp. 171-178.

Samuels, S.E., Environmental strategies for preventing childhood obesity, Acceleration Meeting Memo,
January 8 and 9, 2004, Berkeley Media Studies Group, Berkeley, 2004.

Shanahan, D.F., Fuller, R.A., Bush, R., Lin, B.B., and Gaston, K.J., ‘The Health Benefits of Urban Nature: How
Much Do We Need?’, BioScience, Vol. 65, 2015, pp. 476-485.

Shanahan, D.F., Bush, R., Gaston, K.J., Lin, B.B., Dean, J., Barber, E., Fuller, R.A., ‘Health Benefits from
Nature Experiences Depend on Dose’, Scientific Reports, Vol. 6, Art. no 28551, 2016.

Sinnett, D., Smith, N., and Burgess, S. (Eds.), Handbook on Green Infrastructure: Planning, design and
implementation, Edward Elgar Publishing Ltd, Cheltenham, 2015.

Somarakis, G., Stavros, S., Chrysoulakis, N. (Eds.), ThinkNature Nature-Based Solutions Handbook,
ThinkNature project funded by the EU Horizon 2020 research and innovation programme under grant
agreement No 730338, 2019, doi:10.26225/ jerv-w202.

Tzoulas, K., Korppela, K., Venn, S., Yli-Pelkonen, V., Kaźmierczak, A., Niemela, J., and James, P., ’Promoting
ecosystem and human health in urban areas using Green Infrastructure: A literature review’,
Landscape and urban planning, Vol. 81, No 3, 2007, pp. 167-178.

Uzzell, D., Pol, E., and Badenas, D., ‘Place identification, social cohesion, and environmental sustainability’,
Environment and Behavior, Vol. 34, No 1, 2002, pp. 26-53.

van den Berg, M.M., van Poppel, M., van Kamp, I., Ruijsbroek, A., Triguero-Mas, M., Gidlow, C.,
Nieuwenhuijsen, M.J., Gražulevičiene, R., van Mechelen, W., Kruize, H., and Maas, J., ‘Do physical
activity, social cohesion, and loneliness mediate the association between time spent visiting green
space and mental health?’, Environment and Behavior, Vol. 51, No 2, 2017, pp. 144-166.

Van Den Berg, A.E. and Custers, M.H.G., ‘Gardening Promotes Neuroendocrine and Affective Restoration from
Stress’, Journal of Health Psychology, Vol. 16, No 1, 2011, pp. 3–11.

Varela-Candamio, L., Novo-Corti, I., and García-Álvarez, M.T., ‘The importance of environmental education in
the determinants of green behavior: A meta-analysis approach’, Journal of Cleaner Production, Vol.
170, 2018, pp. 1565-1578.

Vujcic, M., Tomicevic-Dubljevic, J., Grbic, M., Lecic-Tosevski, D., Vukovic, O., and Toskovic, O., ‘Nature-based
solution for improving mental health and well-being in urban areas’, Environmental Research, Vol.
158, 2017, pp. 385-392.

Wendling, L.A., Huovila, A., zu Castell-Rüdenhausen, M., Hukkalainen, and M., Airaksinen, M., ‘Benchmarking
nature-based solution and Smart City assessment schemes against the Sustainable Development Goal
indicator framework’, Frontiers in Environmental Science, Vol. 6, Art. no 69, 2018, 18 pp.

World Health Organisation, Urban green spaces and health, WHO Regional Office for Europe, Copenhagen,
2016.

World Health Organisation, Urban Green Space Interventions and Health: A review of impacts and
effectiveness, WHO Regional Office for Europe, Copenhagen, 2017.

Xiang, P., Wang, Y., and Deng, Q., ‘Inclusive nature-based solutions for urban regeneration in a natural
disaster vulnerability context: A case study of Chongqing, China’, Sustainability, Vol. 9, No 7, 2017,
pp. 1-13.

173
174
Drawing on knowlegde from projects
funded by the European Union

CLEARING HOUSE
Collaborative LEArning in Research, Information-
sharing and Governance on How Urban forests as
nature-based solutions support Sino-European futures
Barcelona (ES) Beijing (CN) Brussels (BE) Krakow (PL) Leipzig (DE)
Gelsenkirchen (DE) Hangzhou (CN) Hong Kong - Guangzhou - Shenzhen (CN)
Huaibei (CN) Xiamen (CN)

CLEARING HOUSE addresses a global challenge that unites European and Chinese cities in their
quest to develop more resilient cities and liveable societies. Our main focus is on tree-based green
infrastructure which is the basis for “urban forests as nature-based solutions”. Urban Forests
as nature-based solutions (UF-NBS) are nature-based solutions that build on tree-based urban
ecosystems to address societal challenges, simultaneously providing ecosystem services for hu-
man well-being and biodiversity benefits. UF-NBS include peri-urban and urban forests, forested
parks, small woods in urban areas, and trees in public and private spaces. CLEARING HOUSE will
analyse and develop the potential of UF-NBS– across China and Europe – in order to enhance the
resilience of cities facing major ecological, socio-economic, and human wellbeing challenges.

Image: Tongzhou urban forest - Photo © Beijing Forestry Bureau


SCOPE Trees are the solution - Implementing and managing tree-based
ecosystems as nature-based solutions for sustainable urbanization

Approach to Impact Assessment Main Challenges addressed

CLEARING HOUSE is focusing on the impact created


by tree-based ecosystems in an urban and peri-ur-
1. Climate Resilience
ban context. CLEARING HOUSE is using innovative
approaches and innovative tools to assess the im- 2. Water Management
pact of trees, woods and forests to the urban en-
vironment, based on a holistic and interdisciplinary 3. Natural and Climate Hazards
analytical framework. CLEARING HOUSE is develo-
ping a citizen science UF-NBS monitoring tool, that 4. Green Space Management
citizens can use to map, asses and monitor UF-NBS
and their socio-ecological impacts. A benchmar- 5. Biodiversity
king tool will allow to compare UF-NBS in different
settings, and will be used as a quick scan to asses 6. Air Quality
UF-NBS designs. The scenario evaluator will allow
to optimize UF-NBS planning, design and manage- 7. Place Regeneration
ment at the local and regional level.
8. Knowledge and Social Capacity Building
Involved Stakeholders and roles
9. Participatory Planning and Governance
The tools that CLEARING HOUSE is developing are ai-
med towards citizens, local and regional authorities, 10. Social Justice and Social Cohesion
landscape architects, environmental planners, urban
planners, architects, decision-makers, politicians and 11. Health and Wellbeing
natural resource managers. But also experts from pu-
blic health and social work will be able to use our tools, 12. New Economic Opportunities & Green Jobs
to support UF-NBS as a tool to combat social injustice
and health inequality. Citizens and citizen groups will
be able to use the citizen science tool, by mapping
existing resources and having their impacts asses-
Lessons learned
sed by the tool. The scenario evaluator and bench-
marking tools are more focused towards experts, We have learned that people and trees have an
decision-makers, and planners: planners and experts intense relationship. This is helpful to find sup-
can design solutions and create diverging scenarios; port and funding for designing, planting and ma-
decision-makers can make informed decisions based naging trees, woods and forests in and around
on the assessment of the diverging scenarios by the towns and cities. The COVID-19 pandemic – and
scenario evaluator and the benchmarking tool. its resulting restrictions – have showed that ur-
ban green spaces are very important to the pu-
blic, and that they offer space for finding peace
Municipal Administrations and to recover, places for recreation and physical
activity, but also to meet other people in a soci-
Citizen ally distant way.

Planning experts However, tree-based impacts are difficult to as-


sess completely, as they provide a range of be-
Scientists / Academia nefits – but also disbenefits. Disbenefits tend to
be overlooked, which in the end can lead to ten-
sions related to trees in the urban area.
NGOs

Landscape businesses

Schools and kindergartens

politicians
Learn more
Natural resource managers www.clearinghouseproject.eu

The CLEARING HOUSE project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 821242.
Drawing on knowlegde from projects
funded by the European Union

REGREEN
Fostering nature-based solutions for equitable, green
and healthy urban transitions in Europe and China
Aarhus (DK) Paris Region (FR) Velika Gorica (HR)
Beijing (CN) Ningbo (CN) Shanghai (CN)

REGREEN aims to substantially advance evidence and tools by systematically modelling


ecosystem services and biodiversity, and examining synergies and trade-offs between
them. This forms the basis for guiding city authorities in effective planning and imple-
mentation of urban NBS in Europe and China. This includes policy experimental learning,
strategies for depavement, education and citizen science in schools, valuation of benefits
and costs and the development of business models for realising spatially relevant NBS
that provide multiple ecosystem services and wellbeing.

Image: Urban Afforestation in Aarhus ULL - Photo © Aarhus Municipality


SCOPE Multiple NBS examined at fine resolution, and assessed at city scale

Approach to Impact Assessment Main Challenges addressed

By creating new ecosystem service models which


take into account local situations, the project will
1. Climate Resilience
produce new tools and guidance which can be used
in the planning stage of NBS design. A strength of 2. Water Management
the tools is to help city authorities find the optimum
location for any intervention, which can satisfy mul- 3. Natural and Climate Hazards
tiple outcomes. The potential impact of new NBS can
be assessed through scenarios, using the models to 4. Green Space Management
evaluate before and after situations at the planning
stage. The metrics produced by the models allow as- 5. Biodiversity
sessment against physical metrics (amount of noise
mitigated, pollution removed, biodiversity enhanced) 6. Air Quality
and societal, health and economic metrics where ap-
propriate (such as number of people experiencing 7. Place Regeneration
reduced heat-stress, economic value of carbon se-
questered). Metrics of success in the project include 8. Knowledge and Social Capacity Building
the number of new NBS projects where scientific out-
comes from REGREEN have helped inform the design 9. Participatory Planning and Governance
or location at the planning stage.
10. Social Justice and Social Cohesion

Involved Stakeholders and roles 11. Health and Wellbeing

In REGREEN, Urban Living Labs (ULLs) are formed 12. New Economic Opportunities & Green Jobs
by municipal administrations, a regional develop-
ment agency and network of local and regional go-
vernments in Europe and China together with local
universities and SMEs. Together, they work on unco- Lessons learned
vering, mapping and engaging a whole ecosystem of
local stakeholders in order to advance the agenda of Although still at an early stage in the project,
regreening cities. For instance, ULL transition work- there is genuine interest from city authorities in
shops in Europe and China will enable sharing of ex- how REGREEN can help design and optimize lo-
perience between ULLs and with local stakeholders; cations for new planned NBS initiatives. To this
experimental policy learning in the ULLs among the end, REGREEN co-creates with city authorities
public, stakeholders and city authorities aims to nur- comprehensive scenarios of NBS interventions
ture innovative and novel governance approaches to that form the basis for assessing the multiple im-
NBS; and co-creation with children, schoolteachers, pacts of NBS in ecosystem models. Impacts co-
park managers and landscape architect students will ver air pollution, urban heat islands, noise, flood-
enhance children’s play and learning activities. ing, water quality and biodiversity. Valuation of
benefits to society and costs of implementation
and maintenance will further help city authorities
Municipal Administrations prioritise and plan NBS interventions.

Citizen

Planning experts

Scientists / Academia

NGOs

Green businesses
Learn more
Schools and kindergartens www.regreen-project.eu

The REGREEN project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 821016
5 APPLICATION OF THE NBS IMPACT EVALUATION FRAMEWORK: NBS
PERFORMANCE AND IMPACT EVALUATION CASE STUDIES
Coordinating Lead authors
Dubovik, M., Dumitru, A., Wendling, L.

Lead authors
Briega, P., Capobianco, V., Connop, S., Crespo, L., Fermoso, J., Giannico, V., Gómez, S.,
González, M., Kakoulaki, G., Kumar, P., Leppänen, S., Marijuan, R., Pablo, S., Pérez, J.A., Pilla,
F., Rinta-Hiiro, V., Riquelme, H., Sánchez, E., Sánchez, I., Sánchez, J.C., Sánchez, R., San
José, E., Sanz, J.M., Sanz, N., Serramia, J., Spano, G., Särkilahti, M., Tomé-Lourido, D., van
de Sijpe, K., Verdugo, F., Villazán, A., Vos, P., Zulian, G.

Contributing authors
Allaert, K., Almenar, J.B., Arnbjerg-Nielsen, K., Baldacchini, C., Basco, L., Beaujouan, V.,
Benoit, G., Bockarjova, M., Bonelli, S., Bouzouidja, R., Butlin, T., Calatrava, J., Calfapietra, C.,
Cannavo, P., Caroppi, G., Chancibault, K., Cioffi, M., Dadvand, P., de Bellis, Y., de Keijzer, C.,
de la Hera, A., Decker, S., Djordjevic, S., Dushkova, D., Faneca, M., Fatima, Z.,; Ferracini, C.,
Fleury, G., García, I., García‐Alcaraz, M., Gerundo, C., Gil-Roldán, E., Giordano, R., Giugni, M.,
Gonzalez-Ollauri, A., Guidolotti, G., Haase, D., Heredida, J., Hermawan, T., Herranz-Pascual,
K., Hölscher, K., Jermakka, J., Kiss, M., Kraus, F., Körmöndi, B., Laikari, A., Laille, P., Lemée,
C., Llorente, M., Lodder, M., Lourido, D.T., Macsinga, I., Manzano, M., Martelli, F., Martins, R.,
Mayor, B., McKnight, U., Mendizabal, M., Mendonça, R., Mickovski, S.B., Nash, C., Nadim, F.,
Nolan, P., Oen, A., Olsson, P., Olver, C., Paradiso, F., Petucco, C., Pisani, N., Piton, G.,
Pugliese, F., Rasmussen, M., Munro, K., Reich, E., Reichborn-Kjennerud, K., Renaud, F.,
Rhodes, M.L., Robles, V., Rodriguez, F., Roebeling, P., Ruangpan, L., Rugani, B., Rödl, A.,
Sánchez Torres, A., Sanesi, G., Scharf, B., Silvestri, F., Skodra, J., Stanganelli, M., Szkordilisz,
F., Tacnet, J.-M., Vay, L., Vella, S., Vercelli, M., Vojinovic, Z., Werner, A., Wheeler, B., Young,
C., Zorita, S., zu-Castell Rüdenhausen, M.

179
Summary

What is this chapter about?

Selecting appropriate indicators of NBS performance and impact can be


challenging, and is context-dependent. In this chapter, we present case studies
from a variety of NBS demonstrations across Europe and Asia that illustrate the
application of the NBS indicators and methods presented in Chapter 4 and
thoroughly described in Evaluating the Impact of Nature-Based Solutions:
Appendix of Methods. Each case study presents a brief NBS description, reasons
for the selection of specific indicators for that particular NBS and a brief overview
of the ways the indicators are applied and/or monitored. The case studies
describe the stakeholders involved in co-design and co-monitoring of NBS and
discuss the barriers and lessons learned during or after the process. Each case
study provides key references for further reading.

The case studies in this chapter focus on the selection of recommended indicators
for NBS performance and impact, which are generally of primary importance
when creating NBS monitoring and evaluation plans. The case studies further
demonstrate how and why additional indicators can be selected to reflect
particular objectives of projects and local challenges.

How can I use this chapter in my work with NBS?

The examples of indicator application illustrate the practice of selecting the


appropriate indicators from the pool of indicators presented in Chapter 4. This
information will aid in understanding why and how to select indicators for
evaluating NBS performance and impact.

Information from the case studies presented in Chapter 5 can be used to support
planning, indicator selection, execution and monitoring of NBS.

When should I use this knowledge in my work with NBS?

We recommend consulting the case studies during the early stages of NBS
planning and deployment, and well before selecting indicators and establishing
NBS monitoring.

How does this chapter link with the other parts of the handbook?

Chapter 5 complements the presentation of NBS indicators (Chapter 4 and


Appendix of Methods) by presenting explicit examples tied to concrete NBS
actions. This chapter assists in making a selection of the indicators listed under
Chapter 4. It provides insights into NBS monitoring approaches described in
Chapters 2, 3 and 6, and alludes to data generation techniques discussed in
Chapter 7.

180
5.1 Introduction to holistic NBS impact assessment using the
framework of recommended indicators

A series of concrete examples of the application of Recommended indicators are


provided here to illustrate the type of narrative it is possible to develop from the
gathered evidence. Specific messages regarding NBS outcomes can be tailored
for different stakeholders, e.g., citizens, investors, policy-makers, etc. The
Recommended indicators illustrated in the following examples reflect the multi-
functionality of NBS and highlight synergies between outcomes in different
societal challenge areas.

For the sake of demonstrating the importance of each individual indicator, the
case studies presented herein describe only the basis for the selection of one, or
in some cases several, either Recommended or Additional indicators (Chapter 4).
This approach was adopted to highlight the importance of the Recommended
indicators as the primary indicators to be addressed when creating NBS
monitoring and evaluation plans, and to emphasise the value of selecting unique
and complementing Additional indicators based on projects’ objectives and the
local challenges NBS aim to address. The case studies were selected per projects’
suggestions given their relative advancement in NBS and their monitoring
strategy implementation. It should be noted that although the case studies
present indicators associated with a specific impact (e.g., water quality or air
quality), the NBS exhibit a much greater number of impacts and co-benefits (e.g.,
on biodiversity, health and well-being), which must be considered when designing
a monitoring strategy.

It is important to note that selected indicators of NBS impact should capture not
only the range of different NBS co-benefits, but should also shed light on trade-
offs for different social groups and between different challenge areas. For
example, issues of gentrification, social justice and similar should be carefully
considered in order to gain an understanding of both benefits and trade-offs, and
to identify potential issues in order to develop effective mitigation strategies.

This Chapter is presented as a series of case studies related to the selection of


Recommended indicators and Additional indicators. Table 6-1 lists the
Recommended and Additional indicators illustrated in the case studies.

181
Table 5-1. Case studies illustrating the selection of Recommended and Additional indicators.

Challenge Recommended indicator Additional indicator


case study case study

Climate
Carbon storage Urban Heat Island incidence
Resilience

Water quality: total suspended


Water solids (TSS) content;

Management Nitrogen and phosphorus
concentration or load

Natural and
– Flood risk
Climate Hazards

Walkability;
Green Space Annual trend in vegetation cover;
Green space accessibility
Management Nature-based recreation;
Land composition

Biodiversity Number of conservation priority


Green infrastructure connectivity
Enhancement species

Air Quality PM10 and PM2.5 concentrations Trends in NOx and SOx emissions

Knowledge and
Social Capacity
Building for
– Connectedness to nature
Sustainable
Urban
Transformation

Social Justice Perceived social support


and Social –
Cohesion

Health and Level of outdoor physical activity Prevalence, incidence, morbidity


Wellbeing (min/week); of chronic stress;
Level of chronic stress Perceived chronic loneliness
("Perceived stress");
Self-reported general wellbeing

182
5.1.1 Recommended indicators case study from Tampere, Finland

NBS Name and Vuores stormwater management system


Location (incl. retention pond, biofilter, alluvial meadows)
Tampere (Finland)

Brief description of The Vuores district is a new district in the City of Tampere
NBS (Finland), featuring an extensive stormwater management system
(in Virolainen- and Tervaslampi Parks) comprising of several NBS,
including the retention pond, biofilter, and alluvial meadows. The
Vuores catchment drains to the Lake Koipijärvi, so preservation of
the lake water quality was the main driver for creating a
comprehensive urban runoff management (quality and quantity)
system.
Virolainen Park:
– Biofilter (with sand as a filtering media): Treatment of
urban runoff and runoff from a dog park
Tervaslampi Park:
– Retention pond: Treatment (retention and
sedimentation) of urban runoff from new housing area
– Alluvial meadows: Space for retention of the urban runoff
at times of heavy rainfall

Useful links:
https://unalab.eu/en/our-cities/city-tampere
www.tampere.fi/unalab (in Finnish)

Indicators of 3.2 Water quality: total suspended solids (TSS) content


relevance 3.3 Nitrogen and phosphorus concentration or load

Explanation for Due to the densification and urbanisation of the newly built areas,
selection of stormwater quality management was the main priority for the City
Indicators in this of Tampere to prevent the water quality deterioration of the local
case waterbodies. TSS content and nutrient (N and P) concentrations
comprise the critical water quality constituents determining the
urban runoff quality entering the surface waterbodies and their
possible adverse effects on the aquatic environment (e.g.,
eutrophication). The NBS addressing water quality further aid in
delivering a variety of co-benefits, including water quantity
management, enhancement of local biodiversity, and contributing
to increased local environmental awareness.

Description of Multiple NBS across the Vuores district are equipped with the
Indicator online water quality sensors continuously measuring a variety of
Application water quality parameters. Each sensor is capable of measuring the
basic water quality parameters, including nitrate-nitrogen (NO3-N)
concentrations. Subsequently, the sensors calculate total
phosphorus concentration based on the turbidity measurements,
and total nitrogen concentration based on the nitrate-nitrogen
measurements. Manual sampling for TSS content is performed at
regular time intervals.

183
Stakeholders City representatives, citizens, NGOs, public and private sector
involved actors (incl. research organisations), and representatives from
universities

Barriers Barriers to ‘physical’ NBS implementation in Tampere included the


encountered and biofilter space requirements in Virolainen Park. Some residents
lessons learned found the alluvial meadows and wetland vegetation (Figure 5-1)
lacking the aesthetics. However, this was overcome through
awareness raising with the information signs and during the co-
creation workshops.
The stakeholder engagement proved to be successful after a series
of co-creation workshops that resulted in the change of plans for
the Vuores area development, additionally considering local
biodiversity, health and water management aspects (Särkilahti
2019).

Case study authors Maria Dubovik1 ([email protected]), Ville Rinta-Hiiro1, Maarit


Särkilahti2, Salla Leppänen2
1
VTT Technical Research Centre of Finland, Espoo, Finland
2
City of Tampere, Finland

References Särkilahti, M., ‘Co-creating nature based solutions in EU project


demonstration city Tampere’, Rakennustekniikka, 2019. Available from:
https://www.ril.fi/fi/rakennustekniikka/teemat/co-creating-nature-
based-solutions-in-eu-project-demonstration-city-tampere.html

Figure 5-1. Nature-based solutions in Vuores Central Park (© City of Tampere).

184
5.1.2 Recommended indicators case study from Valladolid, Spain

NBS name and Urban carbon sink


location Valladolid Demo Site.
The Urban Carbon Sink is located in the eastern part of
the municipality of Valladolid, in the neighbourhood
known as Los Santos-Pilarica (Sector 50, "Los Santos 2").

Brief description of The Urban Carbon Sink (UCS; Figure 5-2) is conceived as an urban
NBS forest in which species have been selected mainly for their ability
to fix carbon. Therefore it is a nature-based solution for the over-
accumulation of carbon dioxide in cities’ atmosphere.

The design of the UCS is embedded into another projected NBS,


the Floodable Park. It will consist in the installation of urban
woodland (initially planned planting 1,500 trees in a 40,000 m2
surface) with appropriate species adapted to temporary flood
condition and with high capacity of carbon sequestration (Fraxinus
spp., Betula spp., Salix spp., Populus spp., etc.). Trees of this
forest will be allocated in specific arboreal series.

This area will be a new urban carbon sink and will form a new
urban ecosystem to preserve the biodiversity. Likewise, this
woodland will provide biomass to energy use with social and
economic purposes.

Expected impacts: The UCS will be located close to industrial and


traffic areas, which act as a source of carbon dioxide emissions
due to combustion processes. This NBS is proposed to compensate
the emissions of this greenhouse gas, capturing it in the form of
biomass.

In order to achieve this effect, it is necessary to include specific


criteria for taxon selection composition and typology of them
during designing stage of UCS. Likewise, it will be essential to take
into account to establish a management plan (pruning, spacing,
etc.).

Multicriteria species assessment is required, focused on C fixation


capacity, in addition with other aspects, such as native vegetation,
easy management, aesthetics, health, ecological coherence and
integrity criteria. Impacts derived from UCS implementation must
be evaluated on medium-long term, since to C fixation capacity of
the species is highly related to the maturity grade of the taxa.

185
Figure 5-2. Urban Carbon Sink conceptual design (URBAN GreenUP project)

Indicators of 1.1 Total carbon removed or stored in vegetation and soil


relevance per unit area per unit time
Temperature decrease
Heatwave risk
Green space distribution (m2/capita)
Green space distribution (km cycle lane/capita)
7.1 Green space accessibility
Green areas sustainability
Elderly people life quality
9.1 Green infrastructure connectivity
Pollinator species increase

Explanation for This NBS will improve the accessibility to green space value in the
selection of area for the surrounded population, with 40.000 m2 of new
Indicators available green space.

Other indicators that are related with this NBS are those related
with Carbon storage, as it is the main purpose of this NBS.

Description of In this case, the main indicator for impact assessment is 01.01
Indicator and 01.02 and additionally the other ones. This indicator will need
Application an spatial and statistical analysis, following the following algorithm
(Figure 5-3):

186
Figure 5-3. Suggested algorithm for the QGIS process as defined in Deliverable D2.4: Monitoring
Program to Valladolid from the URBAN GreenUP Project.

In this case, “Green infrastructures” is referred to the arriving point and “entryways” to
departure point.

Stakeholders Different municipality areas (at least urbanism, environment and


involved heritage), car park property, construction and gardening
companies, River Duero Basin (it is located in the Esgueva River
bank).

Barriers Main barriers are located in the availability of data required for
encountered and this Indicator.
lessons learned

Case study authors Raúl Sánchez1, Jose Fermoso1, Francisco Verdugo1, Raquel
Marijuan1, Silvia Gómez, María González1, José María Sanz1,
Esther San José1
1
CARTIF Foundation. P.T. Boecillo, 205, 47151, Boecillo, Valladolid, Spain

187
5.1.3 Recommended indicators case study from Guildford, UK

NBS name and Roadside green infrastructure


location Guildford, UK

Figure 5-4. Roadside green infrastructure.

Brief description Roadside Green Infrastructure (Figure 5-4) includes trees,


of NBS hedges, individual shrubs, green walls, and green roofs. The focus
of the iSCAPE pilot in Guildford (UK) was air pollution abatement
and in specific on particulate matter (PM), which is composed of
particles such as black carbon (BC). The pilot focused on near-road
environments, where vegetation can act as a barrier between traffic
emissions and pedestrians (figure below), by collecting pollutants
and/or redirecting the flow of polluted air (Abhijith et al., 2017;
Kumar et al., 2019; Riondato et al., 2020; Tiwari et al., 2019). This
study performed as part of iSCAPE (GA nº 689954) pioneered the
adoption of this kind of nature based solution as a passive control
system for roadside pollution in urban street canyon and open road
settings.

The pilot assessed through monitoring and modelling different


combinations of trees, hedges and individual shrubs to assess their
performances in urban street canyon and open road settings in
terms of abatement of road traffic particulate matter (PM).

Project results show that green barriers can produce a reduction of


concentration of Black Carbon up to 52%, PM1 up to 31%, PM2.5 up
to 17%, PM10 up to 15%.

A series of design parameters were also created for both urban


street canyon and open road settings to help planners in the
effective deployment of this kind of air pollution abatement
intervention (Kumar et al., 2019):

Considerations for urban street canyon green infrastructure


Design parameter Considerations
Location If the prime objective is to reduce exposure for pedestrians or
cyclists, hedges should be planted close to the road, between the

188
road and footpath/bike path. Green walls can be constructed on
the pillars of flyovers, retaining walls and other boundary walls.
Selection of In deep street canyons, no forms of vegetation except green
vegetation walls are recommended. In mid-depth street canyons (Table 4),
shrubs or hedges and green walls can be planted, but trees are
not recommended. Large, dense trees should be avoided in all
street canyons, but smaller or lighter-crowned trees may be
planted in shallow street canyons.
Spacing Continuous hedges (with no gaps or spacing) provide a better
reduction in exposure for pedestrians and cyclists. If trees are to
be planted (shallow canyons only), they should be spaced
generously apart from one another.
Height For hedges, a height of around 2m is recommended.
Thickness For hedges, a thickness of 1.5m or more is recommended.
Density In street canyons, a higher density for hedges and lower density
for trees is recommended.

Considerations for open road green infrastructure


Design parameter Considerations
Location Hedgerows should be planted between the road and walkways or
dwellings and in front of trees (if present); this configuration offers
the maximum reduction of exposure.
Spacing Barriers with no gaps provide better downwind exposure reduction.
Height Where possible, it is recommended that the combined hedge-tree
barrier or green wall has a height of 5m or more. Vegetation
barriers with greater height result in increased pedestrian-side
pollutant reductions. A minimum height of 1.5m is recommended.
Thickness The vegetation should be as thick as possible; thicker vegetation
barriers offer greater exposure reduction. If possible, a thickness
of more than 5m is recommended.
Density High-density vegetation barriers are generally better for reducing
exposure levels downwind.

Indicators of Monetary values: value of air pollution reduction; total monetary


relevance value of urban forests including air quality, run-off mitigation,
energy savings, and increase in property values.
11.1 Air quality parameters (Particulate Matter)†
Concentration of particulate matter (PM2.5 and PM10) at respiration
height along roadways and streets.

Contributes to evaluating indicators 12.7/12.8

Explanation for In future, if this NBS is widely installed it can be used recommended
selection of indicators for Air Quality challenges (Figure 5-5). Recommended
Indicators indicators have a scale of measurement from district to region and
they have not sensibility enough to study the impact of this NBS.
Therefore, in the meantime it is needed additional indicators to
assess the impact on air pollutants emission reduction with
indicators such as the ones mentioned before.

189
Description of In this case, the main indicators for impact assessment is 6.11 and
Indicator 6.13. 6.11 implies the installation of sensors for continuous
Application monitoring of PM on the two sides of the deployed green barrier
NBS.

It is also recommended to complement the monitoring campaign


with modelling to account for the impact of local climate.

Stakeholders A wide range of stakeholders including local authorities, academia


involved and local community which were involved in co-design and co-
monitoring activities.

Barriers The main challenge was the initial engagement of the stakeholders
encountered and for the co-design and co-monitoring activities part of the Living Lab
lessons learned framework embraced by iSCAPE. The development of a solid
strategy resulted in a very high engagement of the stakeholders in
this pilot, which allowed to produce the adequate bottom-up support
to push the findings from the pilot into policy within the lifetime of
the project. The findings were endorsed and operationalised as
policy by the Mayor of London
(https://www.london.gov.uk/sites/default/files/green_infrastruture
_air_pollution_may_19.pdf). The pilot clearly demonstrated the
advantages of involving a wide range of stakeholders in the various
stages of the design, development and monitoring of NBS.

It also clearly demonstrated the effectiveness, if appropriately


deployed, of common elements of green infrastructure as passive
control systems for air pollution.

Case study Francesco Pilla1, Prashant Kumar2


authors Spatial Dynamics Lab, University College Dublin, Ireland
1

2
Global Centre for Clean Air Research, University of Surrey, UK

References Abhijith, K.V., Kumar, P., Gallagher, J., McNabola, A., Baldauf, R., Pilla, F.,
Broderick, B., Di Sabatino, S. and Pulvirenti, B., ‘Air pollution abatement
performances of green infrastructure in open road and built-up street
canyon environments–A review’, Atmospheric Environment, Vol. 162,
2017, pp. 71-86.
Kumar, P., Abhijith, K.V. and Barwise, Y., Implementing green infrastructure
for air pollution abatement: General recommendations for management
and plant species selection, 2019.
Riondato, E., Pilla, F., Basu, A.S. and Basu, B., ‘Investigating the effect of
trees on urban quality in Dublin by combining air monitoring with i-Tree
Eco model’, Sustainable Cities and Society, Vol. 61, 2020, p. 102356.
Tiwari, A., Kumar, P., Baldauf, R., Zhang, K.M., Pilla, F., Di Sabatino, S.,
Brattich, E. and Pulvirenti, B., ‘Considerations for evaluating green
infrastructure impacts in microscale and macroscale air pollution
dispersion models’, Science of The Total Environment, Vol. 672, 2019, pp.
410-426.

190
Figure 5-5. An overview of the relationship between air quality and green infrastructure with a matrix
offering local-scale implementation impacts (adapted from Abhijith et al. 2017 and Kumar et al. 2019).

5.1.4 Recommended indicators case study from Genk, Belgium

NBS name and Schansbroek Park (Genk, Belgium)


location

Brief description of Schansbroek Park lies near the source zone of the Stiemerbeek
NBS River and near the coal mine of Waterschei. The park is an
example of NBS for brownfield regeneration (Figure 5-6), as the
area was surrounded by mining activities that were severely
affected natural water management contributing to pollution and
flooding for local residents (Connecting Nature, 2020). The
topography of the area was altered by mining operations and to

191
protect local residences, rainfall and groundwater has had to be
pumped into the Stiemerbeek River. This severe hydrological
impact caused water shortage for natural wetland areas negatively
impacting their biodiversity. Regarding its attractiveness, although
the area has a 16th century defensive structure ‘De Schans’, the
surroundings were unattractive and there was a lack of
recreational infrastructure for visitors, residents and workers
(Green4Grey, 2020).

In view of the state of the area, the Flemish Land Agency (VLM)
together with the city of Genk began a participatory redesign,
where the suggestions made by local citizens (i.e., allotments,
children's play areas, cycling / hiking trails, picnic and meeting
areas) were included in the new plan (Hölscher et al., 2019). In
addition, the redesign involved measures to recreate a ‘wet
ecotope’ by restoring a natural dam and ponds, and transforming
an artificial reservoir from the former mine (Connecting Nature,
2020).

The environmental benefits were powerful, since the biodiversity


and natural conservation of the area were optimized, reducing
flooding and improving water quality. Furthermore, the fact of
regulating the floods provided thermal comfort zones. The benefits
were not only in the environmental dimension but also in public
governance and wellbeing. The new park enhanced the aesthetics
of the area, with new spaces to exercise and meet up. Thus, it
became an attractive space for residents and workers of the
neighbouring Thorpark that allowed citizens to reconnect with
nature, improving physical and mental wellbeing. The fact of
having conducted participatory planning contributed to promoting
social cohesion and environmental stewardship (Connecting
Nature, 2020).

Indicators of 21.1 Level of outdoor physical activity (min/week)


relevance 21.2 Level of chronic stress ("Perceived stress")
21.4 Self-reported general wellbeing
Frequency of social activities in outdoor spaces

Description of The indicators selected to assess the health and wellbeing


Indicator dimension in Schansbroek Park form a coherent framework that
Application allows analysing the NBS effects on citizens.

Starting with the level of outdoor physical activity, defined as self-


reported participation in organized or unorganized sport or
exercise, outdoors, at least once a week (Schipperijn et al., 2013),
is a fundamental indicator to discover if the new redesign of
Schansbroek Park, with its cycling and hiking routes, improves the
healthy habits of users. Knowing the weekly physical activity levels
allow a broad vision of the health and well-being of the area, since
numerous studies in various countries have shown that access to,
and use of, urban green space contributes to increased physical
activity, wellbeing, higher rates of recreational walking and
reduced sedentary time (Almanza et al., 2012; Braubach et al.,
2017; Lachowycz and Jones, 2014; Sallis et al., 2016; Schipperijn
et al., 2013; Sugiyama et al., 2014).

192
Complementarily, the indicator of frequency of social activities in
outdoor spaces, follows the same line, since during the
participatory design process of the new area of Schansbroek,
neighbours and workers suggested including places that allow
social interaction. This interaction is now possible in the park and
represents a great advance in terms of health and well-being
assessment, as green spaces contribute to social cohesion,
fostering social interactions and engagement, promoting a sense
of community (Jennings and Bamkole, 2019; Prezza et al., 2001).

Chronic stress and self-reported wellbeing complete the vision on


the potential impacts of Schansbroek Park can produce in terms of
well-being, specifically mental health. A growing body of empirical
evidence documents the relationship between connection and
contact with green spaces and a greater subjective well-being
(Frumkin et al., 2017; Howell et al., 2011; Howell and Passmore,
2013; Larson et al., 2016; MacKerron and Maurato, 2013;
Pritchard et al., 2020; Wendelboe-Nelson et al., 2019; Zhang et
al., 2014). Contact with natural urban environments can provide
psychological relaxation and stress alleviation, enhancing immune
function, stimulating social cohesion, supporting physical activity,
and reducing exposure to air pollutants, noise and excessive heat
(Braubach et al., 2017; Hartig et al., 2014).

In addition, other indicators were implemented in the field of


Health and Wellbeing, corresponding to indicators: Perceived
restorativeness of NBS and Incidence of obesity among adults, of
the taskforce.

Description of Methodology and data analysis require high expertise in psycho-


Additional Indicator social research but quantitative data collection requires no
Application expertise. During the Connecting Nature project, the data
gathering is conducted after the NBS implementation, but it allows
making comparisons between different areas of the city or
population groups (i.e., users versus no users). Indicator
application was as follows:

Level of outdoor physical activity (min/week)


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)
o T: International Physical Activity Questionnaire (IPAQ)
(International Physical Activity Questionnaires, n.d.).
IPAQ (both long - 27 items, and short form - 7 items)
assesses physical activity undertaken across a
comprehensive set of domains including:
• leisure time physical activity
• domestic and gardening (yard) activities
• work-related physical activity
• transport-related physical activity

Frequency of social activities in outdoor spaces


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)

193
o T: Ad hoc question adapted from Bloesma et al. (2018):
How often do you intentionally go to a green
environment (not your own garden or Schansbroek
Park) for social activities (meeting family or friends,
chatting with neighbours, having a picnic, playing board
games)?

Level of chronic stress ("Perceived stress")


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)
o T: Perceived Stress Scale (Cohen et al. 1983), a self-
report measure intended to capture the degree to which
persons perceive situations in their life as excessively
stressful relative to their ability to cope. Within
Connecting Nature, the PSS-10 version was used
because it was established as the most recommended
form of PSS (as cited in Taylor, 2015, p. 90).

Self-reported general wellbeing


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)
o T: Satisfaction with Life Scale (Diener et al., 1985), a 7-
point scale comprising 5 items that measure individual’s
general satisfaction with own life as a cognitive-
judgmental process (i.e., based on a comparison with a
standard that individual had set for him/herself).

Perceived restorativeness of NBS


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)
o T: Perceived Restorativeness Scale (the short, PRS -
11) (Pasini et al., 2014), a shorter, parallel version of
the Perceived Restorativeness Scale (PRS – 26) (Hartig
et al., 1997), developed to address original
psychometric limitations; PRS is based on the Attention
Restoration Theory (ART; Kaplan, 1995) and its short
version measures an individual’s perception of 4
restorative factors assumed to be present to a greater
or lesser extent in the environment, namely physical
and/or psychological “being-away” from demands on
directed attention, “fascination” a type of attention
assumed to be effortless and without capacity
limitations, the “coherence” and “scope” perceived in an
environment. Participant’s judgments are made on a 0
to 10-point scale.

Incidence of obesity among adults


 Quantitative P: Scale/Scale inventory/Questionnaire (survey
procedure, paper-and-pencil administration, computer-based
administration)
o T: Measurements of Body mass index (BMI). A ratio of
weight to height that is calculated by the following
formula: BMI = weight (kg) ÷ height (m)². For adults,

194
BMIs in the range of 18.5 to 24.9 are considered to be
healthy – and associated with the lowest risk of
mortality and morbidity. Overweight is defined as a BMI
of 25.0 to 29.9; obesity is defined as a BMI of at least
30, with 3 sub-categories (Class I, Class II, and Class
III) that are associated with increasing risk of
cardiovascular disease, type 2 diabetes, and all-cause
mortality (Bhrem and D'Alessio, 2014).

Stakeholders Connecting Nature; Stad Genk; Green4Grey; the Flemish


involved government

Barriers Genk was formerly seen as a Grey City (dominated by hard


encountered and infrastructure), with certain areas of the city disconnected. This
lessons learned made community participation or sense of ownership more difficult
(van de Sijpe et al., 2019). In this sense, community opinion
regarding the site already used was a barrier, local residents
unofficially used the space and there was a lack of interest in
draining their private gardens. However, the biggest barrier was
the cost of the original design. This plan sought to divert pumped
water back to a pond in the nature reserve to raise the water levels
in order to meet ecological goals, but it became cost-prohibitive,
and mono-functional, so the plan had to change.

The lessons learned encompass this change in the redesign of the


area, since less expensive measures were taken but that met the
same objectives, in addition to enhancing the ecological and social
value of the area (van de Sijpe et al., 2019). Active horizontal
cooperation between several departments was needed, as well as
workshops with the residents of the neighbourhood to explain the
project and encourage them to participate in its co-design.
Schansbroek was the first area to be redeveloped in the
Stiemervallei context, so the lessons learned in terms of project
management, stakeholder engagement and citizen communication
will be of great use to scale up in other areas of the city.

Case study author Adina Dumitru1 ([email protected]), David Tomé-Lourido1,


Peter Vos2, Katrien van de Sijpe2
1
University of A Coruña, Spain
2
City of Genk, Belgium

References Almanza, E., Jerrett, M., Dunton, G., Seto, E., and Pentz, M.A., ‘A study of
community design, greenness, and physical activity in children using
satellite, GPS and accelerometer data’, Health and Place, Vol. 18, 2012,
pp. 46-54.
Bloemsma, L.D., Gehring, U., Klompmaker, J.O., Hoek, G., Janssen, N.A.,
Smit, H.A., Vonk, J.M., Brunekreef, B., Lebret, E., and Wijga, A.H.,
‘Green space visits among adolescents: frequency and predictors in the
PIAMA birth cohort study’, Environmental Health Perspectives, Vol. 126,
No 4, 2018, Art. No 047016.
Braubach, M., Egorov, A., Mudu, P., Wolf, T., Ward Thompson, C., and
Martuzzi, M., ’Effects of Urban Green Space on Environmental Health,
Equity and Resilience’, Nature-Based Solutions to Climate Change
Adaptation in Urban Areas: Linkages between Science, Policy and
Practice, SpringerOpen, Cham, 2017, pp. 187-205.
Brehm, B.J. and D’Alessio, D.A., ‘Environmental factors influencing obesity’,
Endotext, MDText.com, Inc., South Dartmouth, 2014, pp. 2000.

195
Cohen, S., Kamarck, T., and Mermelstein, R., ‘A global measure of perceived
stress’, Journal of Health and Social Behavior, Vol. 24, No 4, 1983, pp.
385-396.
Connecting Nature, ‘Schansbroek, Genk – brownfield regeneration’, 2020.
Retrieved from https://connectingnature.eu/oppla-case-study/19379
Diener, E., Emmons, R.A., Larsen, R.J., and Griffin, S., ‘The Satisfaction
With Life Scale’, Journal of Personality Assessment, Vol. 49, No 1, 1985,
pp. 71-75.
Frumkin, H., Bratman, G.N., Breslow, S.J., Cochran, B., Kahn, P.H., Jr.,
Lawler, J.J., Levin, P.S., Tandon, P.S., Varanasi, U., Wolf, K.L., and
Wood, S.A., ‘Nature Contact and Human Health: A Research Agenda’,
Environmental Health Perspectives, Vol. 125, No 7, 2017, pp. 075001.
Green4Grey, ‘Schansbroek (Genk)’, 2020. Retrieved from
https://green4grey.be/en/project-zones/schansbroek-genk
Hartig, T., Korpela, K., Evans, G.W., and Gärling, T., ‘A measure of
restorative quality in environments’, Scandinavian Housing and Planning
Research, Vol. 14, No 4, 1997, pp. 175-194.
Hartig, T., Mitchell, R., de Vries, S., and Frumkin, H., ‘Nature and Health’,
Annual Review of Public Health, Vol. 35, 2014, pp. 207-228.
Howell, A.J., Dopko, R.L., Passmore, H.-A., and Buro, K., ‘Nature
connectedness: Associations with well-being and mindfulness’,
Personality and Individual Differences, Vol. 51, No 2, 2011, pp. 166-171.
Howell, A.J.and Passmore, H.-A., ‘The nature of happiness: Nature affiliation
and mental well-being’, Mental well-being: International contributions to
the study of positive mental health, Springer, New York, pp. 231–257.
Jennings, V. and Bamkole, O., ‘The relationship between social cohesion and
urban green space: An avenue for health promotion’, International
Journal of Environmental Research and Public Health, Vol. 16, No 3,
2019, pp. 452.
Kaplan, S., ‘The Restorative Benefits of Nature: Toward an Integrative
Framework’, Journal of Environmental Psychology, Vol. 15, 1995, pp.
169-182.
Lachowycz, K. and Jones, A., ‘Does walking explain associations between
access to greenspace and lower mortality?’, Social Science and Medicine,
Vol. 107, pp. 9–17.
Larson, L.R., Jennings, V., and Cloutier, S.A., ‘Public Parks and Wellbeing in
Urban Areas of the United States’, PLoS ONE, Vol. 11, No 4, 2016,
e0153211.
MacKerron, G. and Mourato, S., ‘Happiness is greater in natural
environments’, Global Environmental Change, Vol. 23, No 5, 2013, pp.
992-1000.
Pasini, M., Berto, R., Brondino, M., Hall, R., and Ortner, C., ‘How to measure
the restorative quality of environments: The PRS-11’, Procedia-Social
and Behavioral Sciences, Vol. 159, 2014, pp. 293-297.
Prezza, M., Amici, M., Roberti, T., and Tedeschi, G., ‘Sense of community
referred to the whole town: Its relations with neighboring, loneliness, life
satisfaction, and area of residence’, Journal of Community Psychology,
Vol. 29, No 1, 2001, pp. 29-52.
Pritchard, A., Richardson, M., Sheffield, D., and McEwan, K., ‘The
Relationship Between Nature Connectedness and Eudaimonic Well-
Being: A Meta-analysis’, Journal of Happiness Studies, Vol. 21, 2020, pp.
1145-1167.
Sallis, J., Cerin, E., Conway, T., Adams, M., Frank, L., Pratt, M., Salvo, D.,
Schipperijn, J., Smith, G., Cain, K., Davey, R., Kerr, J., Lai, P., Mitáš, J.,
Reis, R., Sarmiento, O., Schofield, G., Troelsen, J., Delfien, V., and
Owen, N., ‘Articles Physical activity in relation to urban environments in
14 cities worldwide: a cross-sectional study’, The Lancet, Vol. 6736,
2016, pp. 348.
Schipperijn, J., Bentsen, P., Troelsen, J., Toftager, M., and Stigsdotter, U.,
‘Associations between physical activity and characteristics of urban
green space’, Urban Forestry and Urban Greening, Vol. 12, 2013, pp.
109–116.
Sugiyama, T., Cerin, E., Owen, N., Oyeyemi, A., Conway, T., Dyck, D.,
Schipperijn, J., Macfarlane, D., Salvo, D., Reis, R., Mitáš, J., Sarmiento,
O., Davey, R., Schofield, G., Orzanco-Garralda, R., and Sallis, J.,
‘Perceived neighbourhood environmental attributes associated with
adults' recreational walking: IPEN Adult study in 12 countries’, Health
and Place, Vol. 28C, 2014, pp. 22-30.

196
Taylor, J.M., ‘Psychometric Analysis of the Ten-Item Perceived Stress Scale’,
Psychological Assessment, Vol. 27, No 1, 2015, pp. 90-101.
van de Sijpe, K., Vos, P., Dick, G., Mowat, L., Dziubala, A., Zwierzchowska,
I., Vandergert, P., Jelliman, S., Connop, S., Nash, C., and González, G.,
Deliverable 9: An interim report on progress towards initiation of city-
wide nature-based solutions exemplars, 2019, CONNECTING Nature,
Grant Agreement number 730222.
Wendelboe-Nelson, C., Kelly, S., Kennedy, M., and Cherrie, J.W., ‘A scoping
review mapping research on green space and associated mental health
benefits’, International Journal of Environmental Research and Public
Health, Vol. 16, No 12, 2019, 2081.
Zhang, J.W., Howell, R.T., and Iyer, R., ‘Engagement with natural beauty
moderates the positive relation between connectedness with nature and
psychological well-being’, Journal of Environmental Psychology, Vol. 38,
2014, pp. 55-63.

Figure 5-6. Schansbroek Park (© Green4Grey).

197
5.2 Case studies illustrating the ‘story of an indicator’ for some of the
additional indicators

The case studies in this section are designed to illustrate the selection and use of
Additional indicators from each of the 12 Challenge areas to examine a specific
aspect of a given NBS. Each case study details the need for use of an Additional
indicator and describes its application and the obtained results (or anticipated
results).

It should be noted that NBS exhibit multiple co-benefits, identification of which is


of outmost importance for evaluating the wider NBS impact. Case studies for
selection of Additional indicators presented herein illustrate the selection of the
unique indicators. They merely serve as examples of versatility of the NBS impact
assessment approach, which can be tailored to local needs and challenges.

5.2.1 Climate Resilience – Urban Heat Island incidence

NBS name and Green façade


location Valladolid Demo Site
Shopping Centre El Corte Inglés,
Calle Constitución, 2. 47001 Valladolid (Spain)

Brief description of Green Facade is a constructive system that allows planting on a


NBS vertical façade. This NBS is built with a substructure and a
waterproof panel. The substructure is affixed to the façade. The
plants grow in a growing medium that is affixed to the panels. The
water of the irrigation system nourishes the plants.

This green wall was built in collaboration with a private company


(El Corte Inglés), and has benefits for every part involved in the
project: the mall, renewing the image of the facade and attracting
new customers, and the city, improving the air quality, climate
regulation, pollination and adding aesthetic values to a grey area
in the city centre of Valladolid. This vertical garden covers an area
of 350 m2 and has more than 14,000 plants (Figure 5-7, Figure
5-8).

Additional 1.5 Heatwave incidence


Indicators of 1.13 Urban Heat Island (UHI) incidence
relevance
1.15 Mean or peak daytime temperature - 1.15.1 Direct
measurement.
6.9 Trends in emissions of NOX and SOX
6.10 Monetary values: value of air pollution reduction; total
monetary value of urban forests including air quality, run-off
mitigation, energy savings, and increase in property values.
6.11 Air quality parameters. NOX and PM.

Explanation for In future, if this NBS is widely installed it can be used


selection of recommended indicators for climate change and Air Quality

198
Additional challenge. Recommended indicators have a scale of measurement
Indicators from district to region and they have not sensibility enough to
study the impact of this NBS. Therefore, in the meantime it is
needed additional indicators to assess the impact on air pollutants
emission reduction with indicators such as the ones mentioned
before.

Description of In this case, the main indicator for impact assessment is 1.5 and
Additional Indicator 1.15 (1.15.1) and additionally the other ones. 1.15 implies the
Application installation of several equipment for continuous monitoring of
temperature and humidity in the green façade location and
reference areas.

Stakeholders Different municipality areas (at least urbanism, environment and


involved heritage), shopping centre company (El Corte Inglés),
construction companies.

Barriers Regarding the NBS implementation, the main barriers were


encountered and administrative and economic. The green façade was installed in a
lessons learned commercial private building in a relevant area of the city. URBAN
GreenUP joined the efforts of the El Corte Inglés technical team,
different areas of the Valladolid city council and the technical
experts of the Project leaded by SingularGreen. After more than 1
year of discussions, it was decided to separate into two
interventions: A structure to support the NBS and the vertical
garden itself. The structure was attached to the existing wall and
it was designed and constructed by El Corte Inglés. Then, Green
Facade was manage with local and EU funds.

Case study authors Jordi Serramia1, Hugo Riquelme1, Patricia Briega1, Alicia Villazán2,
Isabel Sánchez2, Elena Sánchez2, Juan Carlos Sánchez3, Raúl
Sánchez4, Jose Fermoso4, Raquel Marijuan4, Silvia Gómez4, María
González4, José María Sanz4, Esther San José4
1
SingularGreen S.L. C/ Francisco Carratalá Cernuda, 34 Bajo, 03010,
Alicante, Spain
2
VALLADOLID City Council. Plaza Mayor 1, 47001, Valladolid, Spain
3
Tierra Ingeniería S.L. C/ Copenhague, 6, 28230, Las Rozas, Spain
4
CARTIF Foundation. P.T. Boecillo, 205, 47151, Boecillo, Valladolid, Spain

Figure 5-7. The green façade at El Corte Inglés, Valladolid.

199
Figure 5-8. URBAN GreenUP Project: Green Façade construction details (© SingularGreen).

200
5.2.2 Natural and Climate Hazards – Flood risk

NBS Name and Green barrier


Location Gudbrandsdalen Valley, Norway

Brief description of A receded green flood barrier located at Jorekstad in Lillehammer


NBS municipality (Figure 5-9) is proposed to reduce the risk of floods
due to snow melting and extreme rainfall. The NBS consists of
removing the existing flood protection along a section of the
riverbank, and building a new flood barrier, using only natural and
local materials, further upland of the riverbanks. This will provide
space for the river during periods of flooding and improve the
capacity for upstream flood levels, as well as contribute positively
to the flood plain ecosystem. For more information see:
https://phusicos.eu/case_study/valley-of-gudbrandsdalen-
norway/

Additional Risk reduction:


Indicators of 6.13.1 Urban /Residential Areas
relevance 6.13.2 Productive Areas (Agriculture, Grazing, Industries)
6.15.1 Inhabitants
6.15.3 Other People (Workers, Tourists, Homeless)
6.15.4 Elderly, children, disabled
6.16.1 Population
6.17.1 Housing
6.17.2 Agricultural and Industrial Buildings
6.18.1 Roads
6.18.2 Transportation Infrastructures and Lifelines
6.18.3 Lifelines (Water main, Sewerage, Pipeline, etc.)
6.19.1 Buildings
6.22 Flooded Area
6.24 Peak Flow
24.24 Economic Value of the Productive Activities Vulnerable to
Risk (i.e. Economic Value of the Fields, Workers No.)

Technical and feasibility aspects:


14.22 Material used coherence
24.5 Initial costs
24.6 Maintenance costs
24.7 Replacement costs
24.8 Avoided costs
24.9 Payback Period

Environment and ecosystem:


4.48 Physical parameters
4.48 Chemical Pollution Parameters
4.23 Water Storage Capacity Enhancement
6.41 Total Predicted Soil Loss (RUSLE)
10.22 Typical Vegetation Species Cover
10.3.1 Abundance of Ecotones/Shannon Diversity

201
10.25.1 Diversity of Functional Groups (Plant Functional Diversity)
10.25.2 Diversity of Functional Groups (Animal Functional
Diversity)
10.7.1 Sites of Community Importance (SCI) And Special
Protection Areas (SPA)

Society:
8.31.2 Number of Visitors in New Recreational Areas
Different Activities Allowed in New Recreational Areas
8.35.1 New Pedestrian, Cycling and Horse Paths
23.2 Rate of Increase in Properties Incomes
18.1.1 Citizen Involved
18.1.2 Stakeholders Involved
17.3 Public-Private Partnership Activated
17.4 Policies Set Up to Promote NBS
14.7 Social Active Associations
14.17 Natural and Cultural Sites, Made Available
14.25 Viewshed
14.26 Scenic Sites and Landmark Created

Local economy:
24.18 Jobs Created in The Nature-Based Sector
24.19 Jobs Created in The Nature-Based Solution Construction and
Maintenance
24.17 Gross Profit from Nature-Based Tourism
24.15 Touristic Activeness Enhancing
24.33 New Areas Made Available for Traditional Activities
(Agriculture, Livestock, Fishing, ...)

Explanation for The indicators tailored to this case study encompass a total of 47
selection of indicators. The indicators are aggregated to provide information
Additional about the NBS with respect to five ambits: 1) Risk reduction, 2)
Indicators Technical and feasibility aspects, 3) Environment and ecosystem,
4) Effects on the society, and 5) Effects on local economy. These
five ambits form the basis of the NBS assessment framework
developed in the PHUSICOS project (www.phusicos.eu).

Description of Quantitative, risk-related indicators include Peak Flow volume,


Additional Indicator Flooded Area – calculated through hydraulic modelling – and
Application Exposed residential and productive areas, obtained by GIS
mapping. Ecosystem indicators are aimed to assess both the
effects on water quality, such as the Change in physical and
chemical water parameters, and water quantity, such as the Total
predicted soil loss (RUSLE), or enhanced Water storage capacity.
Indicators for assessing the improved value of the forested
floodplain include Typical vegetation species cover, and Diversity
in plant and animal functional groups. Societal-related indicators
include the Number of visitors in the new recreational areas and
New pedestrian/cycling paths, whilst the Number of jobs created
in the nature-based sector is one of the economy-related
indicators.

202
Stakeholders Innlandet County Administration, Lillehammer municipality,
involved Private land owners, Local farmers' association, Norges
Naturvernforbund (Friends of the Earth Norway, an environmental
and nature protection NGO)

Barriers Barriers encountered:


encountered and The tendering process for procurement of goods and services is
lessons learned often not straightforward, there are complaints from bidders who
were not selected, etc.
Local politics and bureaucracy; revision of land use plans, local
elections, etc.
Land owners resisting use of their land, for various reasons, e.g.
o Loss of agricultural land
o General scepticism to NBS, or lack of knowledge
o Economic reasons; want land compensation, lose extra
income from gravel out-take
Lessons learned:
• Plan well ahead. Getting plans through to practical implementation
takes more time than one possibly could think of.
• Bring stakeholders into the process as early as possible, if
possible from scratch; co-creation and co-design of the
measures establishes ‘ownership’ and increases enthusiasm.
• Use their local knowledge wherever possible and show
appreciation.
• Identify potentially ‘problematic’ stakeholders and plan
strategies to handle these.
• If at all possible, choose public land for your NBSs.
• Identify individuals who can be good ambassadors for the
project and work closely with them.
• Procurement can be time consuming. Be as detailed as
possible in the tender documents. Complaints will lead to
serious delays.

Case study author Vittoria Capobianco ([email protected])


Norwegian Geotechnical Institute, Norway

Reference https://phusicos.eu/case_study/valley-of-gudbrandsdalen-norway

203
Figure 5-9. Aerial photo of the area with the location of the existing flood barrier and the new flood barrier
(top); visualization of the area with the potential multiple actions that can be supported by the flood barrier
(by Agence Ter, bottom).

204
5.2.3 Green Space Management – Walkability

NBS Name and Living Lab districts


Location Turin (Italy), Zagreb (Croatia),Dortmund (Germany),
Ningbo (China)

Brief description of During the proGIreg project, this indicator will be calculated for
NBS the Living (LL) district and for the entire city area in each Front-
Runner City (FRC).

Additional 8.37 Walkability


Indicators of
relevance

Explanation for The Walkability index express the likelihood that a particular area
selection of may be covered by walking. It provides additional information on
Additional the urban structure of a city and, in turn, individual districts.
Indicators Additionally, it can be of useful in assess the effects of Land use
changes (pre/post intervention)

Description of The Walkability index is a GIS derived raster image, function of


Additional Indicator connectivity, accessibility and perceived pleasantness with values
Application ranging from 0 to 1 where 1 indicates the most walkable area
(e.g., a park with pedestrian lanes well connected to city hot spots
like residential and working areas) and 0 indicates the least
walkable area (e.g., a major urban road) (Figure 5-10).

The calculation of the Walkability index requires the following


data:
o Pop Density map
o Road Network
o Public Transit (including stops and routes)
o Land Use and zoning: residential, commercial and office,
industrial, institutional (e.g., schools, libraries,
kindergartens), green/park area, and water and wetland
o Digital elevation model

205
Figure 5-10. Example of walkability index (city of Zagreb – preliminary
results by Vincenzo Giannico, University of Bari).

Stakeholders Civil local authorities for data collection during baseline have been
involved involved

Barriers The walkability index is a derived metric that requires a large


encountered and number of input data. This characteristic leads to two major
lessons learned issues: (1) data availability and (2) data harmonization across the
civil local authorities involved.

To date, only two of the four FRCs (i.e., Zagreb and Dortmund)
sent us the requested data. Additionally, of the received data, only
the files received by the city of Zagreb were actually usable as the
rest of the files were not compliant with the model request and
thus were not useful. However, the problem was discussed with
the local authorities of Dortmund, and they assured that the data
will be provided in the correct data type within a short period of
time. The city of Turin, similarly, is committed to provide the data
as soon as possible.

Another issue concerns the harmonization of data across cities.


Given the nature of the input data involved in the calculation of
the Walkability index, it has been found to be difficult to obtain
data acquired in the same year across cities. For example, the
Land Use map provided by city of Zagreb is from 2012 while the
city of Dortmund provided a Land Use map generated in the first
decade of the 2000s. Land Use maps, in particular, are usually
developed on a multiyear basis by local authorities, as the changes
in land use occurring yearly, especially in European cities, are

206
often limited. As a consequence, we will be unable to calculate a
yearly walkability index, as expected initially, but rather one
walkability index before the initiation of the project and, depending
on the availability of the data, another walkability index at the end
of the project.

Lesson learned:
o Data collection can vary across cities and constant
interaction with local authorities is needed.
o Given the nature of the input data, calculating a yearly
walkability index is not feasible.
o Two Walkability index (pre/post intervention) would be
calculated on the basis of the availability of the data.

Case study author Vincenzo Giannico ([email protected])


University of Bari, Italy

References Fan, P., Xu, L., Yue, W., and Chen, J., ‘Accessibility of public urban green
space in an urban periphery: The case of Shanghai’, Landscape and
Urban Planning, Vol. 165, 2017, pp. 177-192.

5.2.4 Green Space Management – Annual Trend in vegetation cover

NBS name and location This indicator is part of a framework applied at European level to
map and assess urban ecosystems condition and ecosystem
services

Brief description of The Green Space Management – Annual Trend in vegetation cover
NBS indicator was implemented to assess changes in vegetation cover
within the Urban Green Spaces (NBS Type 3) in 700 European
Functional Urban Areas (FUAs; Figure 5-11) as part of the Mapping
and Assessment of Ecosystems and their Services (MAES)
initiative:
https://ec.europa.eu/environment/nature/knowledge/ecosystem_
assessment/index_en.htm

207
Figure 5-11. Distribution of European functional urban areas (FUAs; (EU 28 + Norway and Switzerland)
(source: Maes et al., 2020, Chapter 3.1: Urban Ecosystems).

Additional Indicators At European level the following indicators have been implemented:
of relevance
7.1 Green spaces Accessibility
7.2 Share of green urban areas
8.1 Ecosystem services provision (flood control, nature-based
recreation, pollination)
8.2 Annual trend in vegetation cover by urban green
infrastructure
8.31.1 ESTIMAP nature-based recreation
8.38 Land composition
8.39 Land use change and green space configuration
8.40 Soil sealing

Explanation for We defined Urban Green Spaces in European cities according to the
selection of EU GI Strategy (EC, 2013), as “a strategically planned network of
Additional Indicators natural and semi-natural areas with other environmental features
designed and managed to deliver a wide range of ecosystem
services” (EC, 2013). We carried out the analysis including all
natural and semi-natural areas together with all private and public
green spaces within the core cities and the commuting zones.

The capacity of Green Spaces to provide ecosystem services is


linked to the quality and extent of vegetation cover. This indicator

208
examines how and in which direction vegetation cover changed
between 1996 and 2018. Trend detection in Normalized Difference
Vegetation Index (NDVI) time series can help to identify and
quantify recent changes in ecosystem properties.
Description of Figure 5-12 shows the steps needed to derive the indicator.
Additional Indicator
Application

Figure 5-12. Suggested algorithm for the process (source: Maes et al., 2020, Chapter 3.1: Urban
Ecosystems, Factsheet 3_1_109).

A. Data were physically downloaded from Google earth engine (GEE)


B. From the original maps the Urban green Infrastructure (UGI) mask was created:
o B.1. areas where at least once between 1996 and 2018 the highest-NDVI was
greater than 0.4.
C. The Trend analysis employed a non-parametric approach, namely the Theil–Sen
regression. The slopes of the regression approach were tested for their statistical
significance using the p-value of the Mann–Kendall 50 test for slopes (Corbane et al.,
2018; Forkel et al., 2013; Jin et al., 2019; Novillo et al., 2019; Teferi, et al., 2015;
Wang et al., 2018;).
o C.1 Only pixels where the p-value (Mann–Kendall) was less than 0.05 (95%
confidence interval) have been considered to have a significant medium-term trend
and used as a mask to extract all the indicators.
o C.2 we reported the average greenest value in 2010 as reference value.
o C.3 From the Theil–Sen positive or negative slope we extracted the Delta Greenest,
which represent the change direction over the 22 years of analysis.
o C.4 To make the interpretation easier the annual trends were reported in terms of
percentage of change per decade (using the equation proposed by Teferi et al.,
2015) .
o C.5 The TS-Slope was reclassified in 5 classes representing key gradual to abrupt
change types. They were defined using the minimum measurable change (+-0.001)

50
Mann–Kendall is a temporal trend estimator that is more robust than the least-squares slope
because it is much less sensitive to outliers and skewed data (https://clarklabs.org/terrset/).

209
as thresholds for areas with no changes (Guan et al., 2018; Jin et al., 2019; Verbyla,
2008).
D. CLC map was reclassified using the land mosaic model in Densely built up and interface
zone
o Indicators (C1-C2-C3-C4-C5) were extracted in Core cities and Commuting zone
within Densely built up and interface zone only for significant pixels of UGI.
Spatially explicit data are available for the 700 FUA. The indicator could be used at a city level
to study vegetation development within urban parks.

Figure 5-13 shows the percentage of change per decade in vegetation cover. 26% of European
cities present a downward trend, meaning that there is a tendency to loose vegetation. The
balance between abrupt changes (Figure 5-14) confirms the trend.

Figure 5-13. Trends in vegetation cover (%


change/decade), within densely built areas in core
cities. The pie chart shows the proportion of cities for
each category (source: Maes et al., 2020, Chapter 3.1:
Urban Ecosystems).

Figure 5-14 shows the difference between major


greening and major browning in densely built
areas of core cities. It represents a
“compensation indicator”, if it is positive the
upward trend was higher than the downward
trend and greening areas compensated the loss
of green spaces. If it is negative, the land

210
development pattern did not include any solution to compensate the green loss. This indicator
provide insights at urban/regional/national level about the compensation policies taken to
avoid damages created by land take, soil sealing or climate change.

Figure 5-14. Balance between abrupt greening and


browning changes within densely built areas in core
cities. The pie chart shows the proportion of cities for
each category (source: Maes et al., 2020, Chapter
3.1: Urban Ecosystems).

Stakeholders MAES represents the core activity of Action 5 – Target 2 of the EU


involved Biodiversity strategy to 2020. The all process, started in 2013
involved EU Member States, The Commission (DG ENV, DG-JRC),
The European Environmental Agency (EEA) and several other
stakeholders.
Specifically a workshop, held in Brussels in June 2019, provided the
opportunity for stakeholders to engage in the first EU wide
ecosystem assessment.

211
Barriers encountered Main barriers are linked to: expertise requested for the
and lessons learned implementation of the indicator.
Case study author Grazia Zulian ([email protected])
JRC D3 Land Resources
References Corbane, C., Pesaresi, M., Politis, P., Florczyk, J.A., Melchiorri, M., Freire, S.,
Schiavina, M., Ehrlich, D., Naumann, G., and Kemper T., ‘The grey-green
divide: multi-temporal analysis of greenness across 10,000 urban centres
derived from the Global Human Settlement Layer (GHSL)’, International
Journal of Digital Earth, 2018, pp. 101–118.
EC, ‘Green Infrastructure (GI) — Enhancing Europe’s Natural Capital’,
COM(2013) 249 final, 2013, p. 13.
Forkel, M., Carvalhais, N., Verbesselt, J., Mahecha, M.D., Neigh, C.S.R., and
Reichstein, M., ‘Trend Change detection in NDVI time series: Effects of
inter-annual variability and methodology’, Remote Sensing, Vol. 5, No 5,
2013, pp. 2113–2144.
Jin, J., Gergel, S.E., Lu, Y., Coops, N.C., and Wang, C., ‘Asian Cities are
Greening While Some North American Cities are Browning: Long-Term
Greenspace Patterns in 16 Cities of the Pan-Pacific Region’, Ecosystems,
2019, pp. 383-399.
Maes, J., Teller, A., Erhard, M., Condé, S., Vallecillo, S., Barredo, J.I.,
Paracchini, M.L., Abdul Malak, D., Trombetti, M., Vigiak, O., Zulian, G.,
Addamo, A.M., Grizzetti, B., Somma, F., Hagyo, A., Vogt, P., Polce, C.,
Jones, A., Marin, A.I., Ivits, E., Mauri, A., Rega, C., Czúcz, B., Ceccherini,
G., Pisoni, E., Ceglar, A., De Palma, P., Cerrani, I., Meroni, M., Caudullo,
G., Lugato, E., Vogt, J.V., Spinoni, J., Cammalleri, C., Bastrup-Birk, A.,
San Miguel, J., San Román, S., Kristensen, P., Christiansen, T., Zal, N.,
de Roo, A., Cardoso, A.C., Pistocchi, A., Del Barrio Alvarellos, I., Tsiamis,
K., Gervasini, E., Deriu, I., La Notte, A., Abad Viñas, R., Vizzarri, M.,
Camia, A., Robert, N., Kakoulaki, G., Garcia Bendito, E., Panagos, P.,
Ballabio, C., Scarpa, S., Montanarella, L., Orgiazzi, A., Fernandez Ugalde,
O., and Santos-Martín, F., ‘Mapping and Assessment of Ecosystems and
their Services: An EU ecosystem assessment’, EUR 30161 EN,
Publications Office of the European Union, Ispra, 2020.
Novillo, C., Arrogante-Funes, P., and Romero-Calcerrada, R., ‘Recent NDVI
Trends in Mainland Spain: Land-Cover and Phytoclimatic-Type
Implications’, ISPRS International Journal of Geo-Information, Vol. 8, No
1, 2019, p. 43.
Teferi, E., Uhlenbrook, S., and Bewket, W., ‘Inter-annual and seasonal trends
of vegetation condition in the Upper Blue Nile (Abay) Basin: Dual-scale
time series analysis’, Earth System Dynamics, Vol. 6, No 2, 2015, pp.
617–636.
Wang, J., Zhou, W., Qian, Y., Li, W., and Han, L., ‘Quantifying and
characterizing the dynamics of urban greenspace at the patch level: A
new approach using object-based image analysis’, Remote Sensing of
Environment, Vol. 204, 2018, pp. 94–108.

212
5.2.5 Green Space Management - ESTIMAP nature-based recreation

NBS name and This indicator is part of a framework applied at European level to
location map and assess urban green spaces and ecosystem services.

Brief description of The indicator was implemented to assess the capacity of urban
NBS ecosystems to provide nature based recreation opportunities in
700 European Functional Urban Areas (FUAs; see Figure 5-11 in case
study 5.2.4 Green Space Management – Annual Trend in vegetation
cover). This work was part of the EnRoute project:
https://oppla.eu/groups/enroute
https://publications.jrc.ec.europa.eu/repository/handle/JRC115375

Several cities in EnRoute applied the model at a local scale in close


collaboration with the municipalities:
Poznan: https://oppla.eu/casestudy/19236
Tento: https://oppla.eu/casestudy/19228
Oslo: https://oppla.eu/casestudy/19231

Additional At European level the following indicators have been implemented:


Indicators of 7.1 Green spaces Accessibility
relevance
7.2 Share of green urban areas
8.1 Ecosystem services provision (flood control, nature-based
recreation, pollination)
8.2 Annual trend in vegetation cover in urban green infrastructure
8.31.1 ESTIMAP nature-based recreation
8.38 Land composition
8.39 Land use change and green space configuration
8.40 Soil sealing
Spatially explicit data are available for the 700 FUA.

Explanation for Nature based recreation or “Physical and experiential interactions


selection of with natural environment” (CICES, https://cices.eu/) includes a
Additional wide list of possible experience and activities such as biking;
Indicators boating; climbing; hiking; horseback riding, walk the dog in a nice
area; enjoy a local play ground ; find an urban park nearby.

ESTIMAP nature-based recreation was developed to map the


combination of recreation opportunities available in a given location.
The original model (Liquete et al., 2016; Paracchini et al., 2014;
Vallecillo et al., 2019; Zulian et al., 2013), up to now applied at
European scale, was adapted to fit the urban setting. In previous
applications the approach was used in urban context (Zulian et al.,
2017), but focused only on specific local applications and cities, such
as in Barcelona (Baró et al., 2016) or Trento (Cortinovis, Zulian and
Geneletti, 2018).

Urban ESTIMAP -recreation consists of three basic sections:


o The Recreation Potential (RP), which estimates the
potential capacity of ecosystems to support nature-based
recreational activities. It is based on land suitability for
recreation and a combination of the natural features that
influence recreational opportunity provision (e.g.,
proximity to lakes; viewpoints of geological or
geomorphological interest …)

213
o The Opportunity map (OS) expresses the presence of
facilities to enjoy and reach areas with potential
opportunities.
o The Recreation Opportunity Spectrum map (ROS)
combines the Opportunity map (OS) and the Recreation
Potential (RP).

From a modelling point of view the whole approach is based on


‘Advanced multiple layer Look-up Tables” (LUT) and “proximity”
concepts. Advanced LUT consist of a combination of elements,
scored according to their suitability to provide recreation
opportunities. In this application the scores for each input were
generated from either the literature or expert input (Schröter et al.,
2015). The final outcomes are based on cross tabulation and spatial
composition derived from the overlay of different thematic maps
(Zulian et al., 2017).

Figure 5-16 shows an example of ROS map, applied to the FUA of


Padova (Italy).

Figure 5-17 shows the share of areas with high recreation potential
within European FUAs.

Figure 5-15. The approach for mapping recreation opportunities in cities explained for the functional
urban area of Padova, Italy (source: Maes et al., 2019, Box 2).

214
Figure 5-16. Surface area with high recreation potential in European functional urban areas (FUAs)
(source: Maes et al., 2019).

Stakeholders EnRoute is a project of the European Commission in the framework


involved of the EU Biodiversity Strategy and the Green Infrastructure
Strategy. EnRoute provides scientific knowledge of how urban
ecosystems can support urban planning at different stages of policy
and for various spatial scales and how to help policy-making for
sustainable cities. A key pillar of the project is science-policy
interface. Local stakeholders were involved in all the activities
carried on at a local scale.

Barriers Main barriers are linked to: expertise requested for the
encountered and implementation of the indicator.
lessons learned
Case study author Grazia Zulian1, Georgia Kakoulaki2
1
JRC D3 Land Resources
2
JRC C2

References Cortinovis, C., Zulian, G., and Geneletti, D., ‘Assessing Nature-Based
Recreation to Support Urban Green Infrastructure Planning in Trento
(Italy)’, Land, Vol. 7, No 4, 2018, p. 112.
Liquete, C., Piroddi, C., Macías, D., Druon, J.N., and Zulian, G., ‘Ecosystem
services sustainability in the Mediterranean Sea: Assessment of status and
trends using multiple modelling approaches’, Scientific Reports, Vol. 6,
2016, Art. No 34162.

215
Maes J., Zulian G., Günther S., Thijssen M., and Raynal J., ‘Enhancing
Resilience Of Urban Ecosystems through Green Infrastructure. Final
Report’, Publications Office of the European Union, Luxembourg, 2019.
Paracchini, M.L., Zulian, G., Kopperoinen, L., Maes, J., Schägner, J.P.,
Termansen, M., Zandersen, M., Perez-Soba, M., Scholefield, P.A. and
Bidoglio, G., ‘Mapping cultural ecosystem services: A framework to assess
the potential for outdoor recreation across the EU’, Ecological Indicators,
Vol. 45, 2014, pp. 371–385.
Schröter, M., Remme, R.P., Sumarga, E., Barton, D.N. and Hein, L., ‘Lessons
learned for spatial modelling of ecosystem services in support of ecosystem
accounting’, Ecosystem Services, Vol. 13, 2015, pp. 64–69.
Vallecillo, S., La Notte, A., Zulian, G., Ferrini, S., and Maes, J., ‘Ecosystem
services accounts: Valuing the actual flow of nature-based recreation from
ecosystems to people’, Ecological Modelling, Vol. 392, 2019, pp. 196–211.
Zulian, G., Paracchini, M.L., Maes, J., and Liquete, C., ‘ESTIMAP: Ecosystem
services mapping at European scale’, Publications Office of the European
Union, Luxembourg, 2013.
Zulian, G., Stange, E., Woods, H., Carvalho, L., Dick, J., Andrews, C., Baró,
F., Vizciano, P, Barton, D.N., Nowel, M., Rusch, G.M., Aurunes, P.,
Fernandes, J., Ferraz, D., Ferreira dos Santos, R., Aszalós, R., Arany, I.,
Czúcz, B., Priess, J.A., Hoyer, C., Bürger-Patricio, G., Lapola, D., Mederly,
P., Halabuk, A., Bezak, P., Kopperionen, L., and Viinikka, A., ‘Practical
application of spatial ecosystem service models to aid decision support’,
Ecosystem Services, Vol. 29 C, 2018, pp. 465-480.

216
5.2.6 Green Space Management – Land composition

NBS name and This indicator is part of a framework applied at European level to
location map and assess urban ecosystems condition and ecosystem
services
Brief description of The indicator was implemented to assess Land composition in 700
NBS European Functional Urban Areas (FUAs; see Figure 5-11 in case
study 5.2.4 Green Space Management – Annual Trend in vegetation
cover).

This work was part of the EnRoute project and the MAES initiative.
https://oppla.eu/groups/enroute
https://publications.jrc.ec.europa.eu/repository/handle/JRC115375

Mapping and Assessment of Ecosystems and their Services –


MAES:
https://ec.europa.eu/environment/nature/knowledge/ecosystem_as
sessment/index_en.htm

Additional At European level the following indicators have been implemented:


Indicators of
relevance 7.1 Green spaces Accessibility
7.2 Share of green urban areas
8.1 Ecosystem services provision (flood control, nature-based
recreation, pollination)
8.2 Annual trend in vegetation cover in urban green infrastructure
8.31.1 ESTIMAP nature-based recreation
8.38 Land composition
8.39 Land use change and green space configuration
8.40 Soil sealing

Explanation for Land composition is a measure of the spatial distribution of elements


selection of or components of a landscape. It is used to consider the co-
Additional occurrence of land types within each FUA. It represents the
Indicators arrangements of ecosystem types within and around cities (Figure
5-17).

To quantify land composition we use the Landscape Mosaic (LM),


model available in Guidos tool box
https://forest.jrc.ec.europa.eu/en/activities/lpa/gtb/ (Vogt and
Riitters, 2017).

This indicator is useful to describe the context where NBS are


deployed.

217
Figure 5-17. Land Mosaic maps in Helsinki (FI) and Naples (IT). A = Agriculture; D = Developed; N =
natural; Mix = mixed presence of all land classes (source: Maes et al., 2019).

Description of Spatially explicit data are available for the 700 FUA.
Additional
Indicator In EnRoute the indicator was applied to explore the capacity of
Application urban ecosystems to provide Ecosystem services city types based on
land composition and population density. Urban Atlas
(https://land.copernicus.eu/local/urban-atlas) was used as land
cover dataset.

Figure 5-18. shows EU FUA classified with reference to land


composition, population density and size.

218
Figure 5-18. Spatial distribution of European functional urban areas (FUAs) classified by land
composition, size and population density. The map includes FUAs in Norway and Switzerland (source:
Maes et al., 2019).

Figure 5-19 shows the behaviour of two indicators (8.31.1 ESTIMAP nature based recreation
and 7.2 share of urban green) with respect to the typology of cities. The indicators exhibit a
high variability in average per city type as well as a high variability in the range of values.
This is especially evident for the share of green spaces in core cities.

Figure 5-19. Average and range of the share of FUA with high recreation potential and share of green
spaces per core city (source: Maes et al., 2019).

219
In MAES the indicator was applied to analyse the changes in land composition (Figure 5-20).
Corine land Cover (https://land.copernicus.eu/pan-european/corine-land-cover ) was used as
land cover dataset.

Figure 5-20. FUAs classified in terms of magnitude and direction of change between 2000 and 2018.
(source: Maes et al., 2020, Chapter 3.1: Urban Ecosystems; Factsheet 3.1.107).

Stakeholders EnRoute is a project of the European Commission in the framework


involved of the EU Biodiversity Strategy and the Green Infrastructure
Strategy. EnRoute provides scientific knowledge of how urban
ecosystems can support urban planning at different stages of policy
and for various spatial scales and how to help policy-making for
sustainable cities. A key pillar of the project is science-policy
interface. Local stakeholders were involved in all the activities carried
on at a local scale.

MAES represents the core activity of Action 5 – Target 2 of the EU


Biodiversity strategy to 2020. The all process, started in 2013
involved EU Member States, The Commission (DG ENV, DG-JRC), The
European Environmental Agency (EEA) and several other
stakeholders.

Specifically, a workshop, held in Brussels in June 2019 provided the


opportunity for stakeholders to engage in the first EU wide ecosystem
assessment.

Barriers Main barriers are linked to: expertise requested for the
encountered and implementation of the indicators.
lessons learned
Case study author Grazia Zulian ([email protected])
JRC D3 Land Resources

References Maes, J., Zulian, G., Günther, S., Thijssen, M., and Raynal, J., ‘Enhancing
Resilience Of Urban Ecosystems through Green Infrastructure. Final
Report’, Publications Office of the European Union, Luxembourg, 2019.
Maes, J., Teller, A., Erhard, M., Condé, S., Vallecillo, S., Barredo, J.I.,
Paracchini, M.L., Abdul Malak, D., Trombetti, M., Vigiak, O., Zulian, G.,
Addamo, A.M., Grizzetti, B., Somma, F., Hagyo, A., Vogt, P., Polce, C.,
Jones, A., Marin, A.I., Ivits, E., Mauri, A., Rega, C., Czúcz, B., Ceccherini,
G., Pisoni, E., Ceglar, A., De Palma, P., Cerrani, I., Meroni, M., Caudullo,

220
G., Lugato, E., Vogt, J.V., Spinoni, J., Cammalleri, C., Bastrup-Birk, A., San
Miguel, J., San Román, S., Kristensen, P., Christiansen, T., Zal, N., de Roo,
A., Cardoso, A.C., Pistocchi, A., Del Barrio Alvarellos, I., Tsiamis, K.,
Gervasini, E., Deriu, I., La Notte, A., Abad Viñas, R., Vizzarri, M., Camia,
A., Robert, N., Kakoulaki, G., Garcia Bendito, E., Panagos, P., Ballabio, C.,
Scarpa, S., Montanarella, L., Orgiazzi, A., Fernandez Ugalde, O., and
Santos-Martín, F., ‘Mapping and Assessment of Ecosystems and their
Services: An EU ecosystem assessment’, EUR 30161 EN, Publications Office
of the European Union, Ispra, 2020.
Vogt, P. and Riitters, K., ‘GuidosToolbox: universal digital image object
analysis’, European Journal of Remote Sensing, Vol. 50, No 1, 2017, pp.
352–361.

5.2.7 Biodiversity Enhancement – Number of conservation priority species

NBS Name and Growchapel and Bellahouston Open Spaces sites


Location Glasgow, UK

Brief description of As part of Glasgow City Council’s Open Space Strategy, they are
NBS rolling out a programme of nature-based solutions to provide
targeted multifunctionality to underused open spaces across the
city. The programme empowers NGOs and community groups to
utilise local spaces and deliver permanent and meanwhile uses on
them including the development of nature-based solutions.
Interventions comprise anything from art installations, to pocket
parks and urban grow-your-own spaces (Figure 5-21).
Multifunctionality is at the heart of the design and Connecting
Nature is supporting the out-scaling of the programme through
greater focus on a nature-based solution approach, more support
for NGOs and community groups to deliver sustainable
stewardship plans, and a spatial dataset of ecosystem service
needs across the city to support decision-making in relation to the
design of the underused spaces.

https://connectingnature.eu/glasgow
https://connectingnature.eu/oppla-case-study/19384

Additional 10.16 Number of conservation priority species


Indicators of 7.1 Greenspace accessibility
relevance 9.1 Greenspace connectivity

Explanation for Whilst biodiversity net-gain is a target of Glasgow City Council’s


selection of Open Space Strategy, these projects are typically delivered in
Additional small spaces and do not have the budgets to cover comprehensive
Indicators biodiversity evaluations (e.g., Recommended biodiversity
indicators like species diversity and functional connectivity). As
such, a more targeted biodiversity indicator was needed.
Evaluation of priority species associated with the spaces was seen
as a win-win for the council as, it represented a more focused
evaluation methodology, and it aligned more closely with strategic
objectives of the local authority and existing monitoring
programmes.

221
Description of Before and after priority species evaluation would be carried out
Additional Indicator to assess any impact of the implemented nature-based solution.
Application This would comprise a combination of local record searches and
direct site evaluation.

Stakeholders This evaluation would be carried out in collaboration with other


involved monitoring schemes in the city (e.g., RSPB sparrow monitoring)
and with other departments in within the council (e.g., biodiversity
team).

Barriers Establishing contacts with appropriate departments and


encountered and organisations was a challenge. Also identifying necessary
lessons learned expertise to carry out surveys.

Case study author Stuart Connop ([email protected])


University of East London, UK

References Connecting Nature Environmental Indicators review:


https://connectingnature.eu/nature-based-solution-evaluation-indicators-
environmental-indicators-review

Figure 5-21. Glasgow meanwhile space conversion providing a temporary grow-your-own space for the
local community (© Glasgow City Council).

222
5.2.8 Air Quality – Trends in NOx and SOx emissions

NBS name and location Urban garden biofilter for air pollution
Underground car park in Portugalete Square
Plaza de la Libertad, 5, 47002 Valladolid (Spain)

Brief description of NBS Urban Garden Biofilter is an air filter framed in an urban
garden for the emissions of underground car parks or
other stationary sources of pollutant compounds in urban
environments. This NBS has been firstly prototyped for
URBAN GreenUP Project (GA nº 730426).

The NBS is composed of three main elements, the extractor


system to extract the polluted air from underground car park,
the plenum section to distribute the air under the Biofilter and
the Biofilter itself to clean the air and metabolize pollutants
(Figure 5-22).

It is composed by several layers for support, pollutants


absorption and protection and finally is cover by vegetation.
The absorption/capture of air pollutants is made by the
different layers and the metabolisation of these pollutants is
made by the soil microbiota and the vegetation.

This NBS has been developed by CARTIF in a previous


research project. Project results show that it can be captured
most of NOX and PM (>90%) from indoor air (pollutants
concentration 0.5-1 ppm).

This NBS can be adapted to existing car parks or tunnels or


included in the design of new infrastructures. It can be
created a new line for indoor air extraction and conduct it to
the plenum zone. Then, the air will be cleaned by passing
thought the biofilter materials. Due to the specific design of
the biofilter layers, pressure drop of the filter is very low and
simple extractor fan is used.

Figure 5-22. URBAN GreenUP Project: Biofilter cross section (© CARTIF).

Additional Indicators of 6.9 Trends in emissions NOx and SOx


relevance

223
6.10 Monetary values: value of air pollution reduction; total
monetary value of urban forests including air quality, run-off
mitigation, energy savings, and increase in property values.
6.11 Air quality parameters. NOX and PM.
6.13 Concentration of particulate matter (PM2.5 and PM10) at
respiration height along roadways and streets.

Explanation for In future, if this NBS is widely installed it can be used


selection of Additional recommended indicators for Air Quality challenge.
Indicators Recommended indicators have a scale of measurement from
district to region and they have not sensibility enough to
study the impact of this NBS. Therefore, in the meantime it
is needed additional indicators to assess the impact on air
pollutants emission reduction with indicators such as the ones
mentioned before.

Description of Additional In this case, the main indicator for impact assessment is 6.11
Indicator Application and additionally the other ones. 6.11 implies the installation
of three equipment for continuous monitoring of NO2, O3 and
PM (inside of the car park, next to the biofilter and separated
from the biofilter but in the same square or street).
This indicator is completed with the other in order to value
and compare biofilter impact with other NBS such as tree or
bush lines.

Stakeholders involved Different municipality areas (at least urbanism, environment


and heritage), car park property, construction companies

Barriers encountered The main difficult aspect is found in the design and project
and lessons learned phase for the implementation of this NBS. Impact assessment
can be carried out by using one or several of the indicators
depending on the budget or monitoring tool available.
Indicator 6.11 is highly recommended and monitoring
locations should be done by experts for the first studies
because this is an innovative solution. The implementation of
this NBS is still ongoing so no experience has been collected
from the monitoring. However, when ongoing pilot studies
and field analysis finish, the assessment framework can be
made simpler by using indicators such as 6.9 or 6.13.

Case study authors Raúl Sánchez1, Jose Fermoso1, Francisco Verdugo1, Raquel
Marijuan1, Silvia Gómez, María González1, José María Sanz1,
Esther San José1, Alicia Villazán2, Isabel Sánchez2, Elena
Sánchez2, Natividad Sanz3, José Antonio Pérez4, Laura
Crespo5
1
CARTIF Foundation. P.T. Boecillo, 205, 47151, Boecillo, Valladolid,
Spain
2
VALLADOLID City Council. Plaza Mayor 1, 47001, Valladolid, Spain
3
ISOLUX CORSAN aparcamientos. Plaza Portugalete, s/n, 47002
Valladolid, Spain.
4
CONYTRAIR. Ctra. Cabezón, 6, 47155 Santovenia de Pisuerga,
Valladolid.
5
LAURA CRESPO ARCHITECT, Valladolid, Spain

224
5.2.9 Knowledge and Social Capacity Building for Sustainable Urban Transformation
– Connectedness to nature

NBS Name and Living Lab districts


Location In the cities of Turin (Italy), Zagreb (Croatia) and
Dortmund (Germany)

Brief description of During the proGIreg project (https://progireg.eu/), this indicator


NBS will be assessed on the general population in the Living (LL)
district and 300 in a different, comparable city district (“control
district”) in each European Front-Runner City (FRC).

Additional 16.3 Mindfulness/ Connectedness to nature


Indicators of 22.13 Perceived restorativeness of NBS/ green space
relevance

Explanation for This indicator is widely used in social sciences since it provides a
selection of reliable assessment of the relationship between human being and
Additional the natural environment
Indicators

Description of Connectedness with nature is defined as the sense of oneness to


Additional Indicator nature. This indicator is part of the socio-cultural inclusiveness
Application evaluation as a component of a survey for the assessment health,
social and economic benefits of NBSs. The “Connectedness to
nature scale” (CNS; Mayer, 2004), a validated tool for assessing
this indicator, will involve 300 persons in each district during two
time points, i.e., pre- and post- NBS implementations (after three
years). The scale includes 14 items with a 5-point Likert scale
ranging from “Strongly disagree” to “Strongly agree”.

Stakeholders Civil local authorities and university students for data collection
involved during baseline have been involved

Barriers The three European FRCs followed a standardized procedure for


encountered and recruitment and data collection, in accordance with the proGIreg
lessons learned scientific WP. Despite the support of the scientific WP through
informal exchange of information and formal meetings in order to
implement strategies to reach the target number of completed
questionnaires, the final outcome differed within the FRCs. The city
of Dortmund has collected 140 interviews (48 in the LL and 92 in
the control district), the city of Turin has collected 398 interviews
(221 in the LL and 177 in the control district). Only the city of
Zagreb managed to reach and even exceeded the determined
target number of interviews, previously set at 600 (302 from the
LL and 313 from the control district).

All cities sent a first information letter to the population in order


to invite to participate in our research. In Turin, the invitation
letters were sent a second time. As expected, the response rate
was very variable between cities and was between 15% and 40%.
The information reported by the cities provides useful insights for
future planning of questionnaires, of which Connectedness with

225
nature scale is part. Participants from each FRC complained about
some aspects of the general questionnaire such as the excessive
length and the presence of uncomfortable questions. No
complaints were specifically addressed to the Connectedness with
nature scale.

Lessons learned regards the strategies that each FRC implemented


to overcome the barriers encountered in reaching the target
number of participants, briefly summarized below.
- Application of a door-to-door technique to directly
approach the target population
- Organization of public events in the neighbourhoods
concerned in order to increase the sample size.
- Second sending of invitation letters following the
unsatisfactory response of the population to the first
sending.
- Possibility of hiring specialized personnel to conduct the
survey.

Case study author Giuseppina Spano ([email protected])


University of Bari, Italy

References Mayer, F., ‘The connectedness to nature scale: A measure of individuals’


feeling in community with nature’, Journal of Environmental
Psychology, Vol. 24, 2004, pp. 503-515.

5.2.10 Social Justice and Social Cohesion – Perceived social support

NBS Name and Living Lab districts


Location In the cities of Turin (Italy), Zagreb (Croatia) and
Dortmund (Germany)

Brief description of During the proGIreg project (https://progireg.eu/), this indicator


NBS will be assessed on the general population in the Living (LL) district
and 300 in a different, comparable city district (“control district”)
in each European Front-Runner City (FRC).

Additional 20.4.1 Perception of socially supportive network


Indicators of 20.4.2 Perceived social support
relevance

Explanation for Empirical evidences showed that supportive social groups and
selection of effective and helpful social networks are associated with a good
Additional mental and physical health. This indicator is measured in the
Indicators neighbour-hood context since a perception of high social support
fosters social inclusion and justice.

226
Description of Perceived social support is defined as the perception of various
Additional Indicator ways in which individuals aid others. This indicator is obtained
Application using an 8-point scale on general social support and a 6-point
scale on social support in the neighbourhood.

Stakeholders Civil local authorities and university students for data collection
involved during baseline have been involved

Barriers The three European FRCs followed a standardized procedure for


encountered and recruitment and data collection, in accordance with the proGIreg
lessons learned scientific WP. Despite the support of the scientific WP through
informal exchange of information and formal meetings in order to
implement strategies to reach the target number of completed
questionnaires, the final outcome differed within the FRCs. The city
of Dortmund has collected 140 interviews (48 in the LL and 92 in
the control district), the city of Turin has collected 398 interviews
(221 in the LL and 177 in the control district). Only the city of
Zagreb managed to reach and even exceeded the determined
target number of interviews, previously set at 600 (302 from the
LL and 313 from the control district).

All cities sent a first information letter to the population in order


to invite to participate in our research. In Turin, the invitation
letters were sent a second time. As expected, the response rate
was very variable between cities and was between 15% and 40%.
The information reported by the cities provides useful insights for
future planning of questionnaires, of which the scale on perceived
social support is part. Participants from each FRC complained
about some aspects of the general questionnaire such as the
excessive length and the presence of uncomfortable questions. No
complaints were specifically addressed to the perceived social
support scale.

Lessons learned regards the strategies that each FRC implemented


to overcome the barriers encountered in reaching the target
number of participants, briefly summarized below.
- Application of a door-to-door technique to directly
approach the target population
- Organization of public events in the neighbourhoods
concerned in order to increase the sample size.
- Second sending of invitation letters following the
unsatisfactory response of the population to the first
sending.
- Possibility of hiring specialized personnel to conduct the
survey.

Case study author Giuseppina Spano ([email protected])


University of Bari, Italy

References Pearson, J.E., ‘The definition and measurement of social support’,  Journal
of Counseling and Development, Vol. 64, 1986, p. 390-395.

227
5.2.11 Health and Wellbeing – Prevalence, incidence, and morbidity of chronic stress

NBS Name and Stalled Spaces


Location Glasgow, Scotland

Brief description of Description


NBS Stalled Spaces (Figure 5-23) is a programme launched by Glasgow
City Council to support community groups and local organisations
across the city develop temporary projects on stalled sites or
under-utilised open spaces. In particular, the Stalled Spaces
programme gives local organizations the opportunity to
temporarily use a plot of these spaces in a way which will bring
multiple benefits to the local communities.

Projects supported by the programme deliver a range of initiatives


based on the needs of the community. It means that community
stakeholders decide how to use these spaced and how to adapt
them to cover their needs. Examples of these initiatives are:
growing spaces, pop-up gardens, wildlife areas, urban gyms or
natural play spaces, temporary art in the form of pop-up
sculptures, and spaces for events or exhibitions.

Relevance
The programme was started in 2011 and only in its first five years
has helped deliver over 100 projects that have successfully
brought over 25 ha of vacant, underutilised or stalled sites under
temporary community use.

Additional 22.22 Prevalence, incidence, morbidity of chronic stress


Indicators of Short name: Chronic stress
relevance
Definition: Within Connecting Nature, stress is defined as the
process by which an individual responds psychologically,
physiologically, and often with behaviours, to a situation that
challenges or threatens well-being (Baum et al., 1985 as cited in
Ulrich et al., 1991, p. 202). The psychological component includes
cognitive appraisal of the situation, emotions such as fear, anger,
and sadness, and coping responses (Ulrich et al., 1991).

Explanation for 1. Theoretical pertinence. Two theoretical frameworks that


selection of establish an association between exposition to / engagement
Additional with nature and stress alleviation have been identified:
Indicators Attention Restoration Theory (ART) (Kaplan, 1995) and
Stress Recovery Theory (SRT) (Ulrich et al., 1991).
2. Impact of the health problem. Chronic stress associated
to modern urban lifestyles is a serious health problem with
an increasing incidence around the world. Moreover,
psychological stress is considered as a significant factor in
the onset, course and exacerbation of other chronic diseases
(depression, cardiovascular diseases…) and it has been
related to the higher overall mortality (Cohen et al., 2007;
Hammen, 2005; Klein et al., 2016).

228
3. Appropriateness of the NBS characteristics. The
multiple initiatives launched in the frame of the Stalled
Spaces Programme over the last decade have not only
contributed to regenerate some areas in Glasgow, but also
to revitalize local communities, to reconnect people with
nature, to generate opportunities for social interaction, to
stimulate social cohesion or to support physical activity. Each
of these achievements constitutes mechanisms to alleviate
chronic stress associated to urban lifestyle and needs to be
explored further to understand how they work and how they
could be reinforced to become more effective.
4. Indicator strengths. Chronic stress is considered as a
reliable indicator to assess physical and mental health and
general wellbeing. In addition, it is appropriate to explore
whether the exposition to a NBS contributes to mitigate
stress.

Description of The tool selected and applied by Glasgow to measure the chronic
Additional Indicator stress indicator in the Stalled Spaces programme is the 10-items
Application Perceived Stress Scale (Cohen et al., 1983) included in a survey
with other indicators specifically chosen to assess the multiple
benefits associated to the implementation of this programme. This
scale is a self-report measure that provides psychological
subjective data. In particular, it intends to capture the degree to
which persons perceive situations in their daily life as excessively
stressful in relation to their ability to cope with them.
Methodology and data analysis require high expertise in psycho-
social research but quantitative data collection does not require
expertise.

Stakeholders Glasgow City Council; Connecting Nature partners; Data collection


involved experts (responsible for collecting subjective psychological data)

Barriers Barriers encountered


encountered and Given the complex psychophysiological pathways of stress,
lessons learned measurement is usually approached holistically through collection
of both subjective psychological (i.e., subjective rating scales, self-
report measures) and objective physiological data (most
frequently, salivary analysis due to the validity, reliability and ease
of collection of salivary data). However, collecting biochemical
data for evaluating a NBS is considered as a major challenge by
the majority of cities for two main reasons: (i) data collection and
analysis of biochemical samples require high clinical expertise,
resources and capacities which are frequently difficult to acquire
for cities; (ii) barriers usually encountered during fieldwork
planning -and in particular those related to the recruitment of
participants - for any study increase when clinical procedures are
included in the design. This means that this objective physiological
measure is feasible in the experimental research usually
conducted by academic and health organizations, but not in the
frame of a routine evaluation conducted by cities.
Lessons learned
1. The experience of Glasgow has demonstrated that it is
essential to provide a detailed description of the
characteristics of the NBS under evaluation and, in particular,
of the activities deployed in it (i.e., gardening, urban gyms,

229
play spaces...). The high diversity of uses allocated to the
Stalled Spaces in Glasgow constitutes an unexceptional
opportunity to identify which activities have a most positive
impact in the stress alleviation (i.e., comparing activities that
enhance physical activity with those that promote social
interaction).
2. In order to gain a holistic understanding of the NBS impact
on the physical and mental health, it is also recommended to
measure this indicator in combination with other indicators
that could contribute to enrich data analysis and
interpretation. In particular, it is suggested to also collect
data about place attachment; general wellbeing and
happiness; and depression and anxiety.
3. It is strongly recommended to collect data on symbolic /
affective meanings assigned to NBS using participatory data
collection methods and qualitative techniques. These data
are useful to understand why and how the exposition to, and
the engagement with, the NBS could contribute to alleviate
chronic stress.

Case study authors Adina Dumitru1 ([email protected]), David Tomé-Lourido1,


Susana Pablo1
1
University of A Coruña, Spain

References Cohen, S., Kamarck, T., and Mermelstein, R., ‘A global measure of perceived
stress’, Journal of Health and Social Behavior, Vol. 24, No 4, 1983, pp.
385-396.
Cohen, S., Janicki-Deverts, D., and Miller, G. E., ‘Psychological stress and
disease’, Journal of the American Medical Association, Vol. 298, No 14,
2007, pp. 1685-1687.
Glasgow City Council, Open Space Strategy, 2020.
Hammen, C., ‘Stress and Depression’, Annual Review of Clinical Psychology,
Vol. 1, 2005, pp. 293-319.
Kaplan, S., ‘The Restorative Benefits of Nature: Toward an Integrative
Framework’, Journal of Environmental Psychology, Vol. 15, 1995, pp.
169-182.
Klein, E.M., Brähler, E., Dreier, M., Reinecke, L., Müller, K.W., Schmutzer,
G.G., Wölfling, K., and Beutel, M.E., ‘The German version of the
Perceived Stress Scale – psychometric characteristics in a representative
German community sample’, BMC Psychiatry, Vol. 16, 2016, pp. 1-10.
Ulrich, R.S., Simons, R.F., Losito, B.D., Fiorito, E., Miles, M.A., and Zelson,
M., ‘Stress recovery during exposure to natural and urban
environments’, Journal of Environmental Psychology, Vol. 11, No 3,
1991, pp. 201-230.
White, J.T. and Bunn, C., ‘Growing in Glasgow: Innovative practices and
emerging policy pathways for urban agriculture’, Land Use Policy, Vol.
68, 2017, pp. 334-344.

230
Figure 5-23. Stalled Spaces Programme (© Glasgow City Council).

5.2.12 Health and Wellbeing – Perceived chronic loneliness

NBS Name and Bellahouston Demonstration Garden


Location Glasgow, Scotland

Brief description of Bellahouston Demonstration Garden was established in the city of


NBS Glasgow, providing allotment-style growing spaces to be used by
different charities and educational establishments (Hölscher et al.,
2019; White and Bunn, 2017). The NBS arises from the Allotment
and Neighbourhood and Sustainability strategies, carried out by
the Glasgow City Council, highlighting the restorative and
therapeutic benefits of gardening, due to social interaction in the
community (White and Bunn, 2017).
The objective of this growing space located in the walled Garden
at Bellahouston Park is twofold, on the one hand to provide
healthy and sustainable food to the neighbours, and on the other
hand to create a community space with social and health benefits
for the citizens of Glasgow.

Additional 22.9 Perceived chronic loneliness


Indicators of Within Connecting Nature, this indicator is conceptualized as a
relevance subjective experience of being socially isolated and absent both
relational and collective connectedness (Russell et al., 1980).

231
Explanation for The strategies implemented for the creation of demonstration
selection of gardens and growing spaces in Glasgow seek to promote social
Additional interaction and engaging people who felt isolated from the
Indicators community (White and Bunn, 2017). Social isolation has a lasting
impact on health and wellbeing (e.g., increased levels of stress,
depression, or cardiovascular concerns) (Holt-Lunstad et al.,
2010; Holt-Lunstad et al., 2015; Pantell et al., 2013), while social
cohesion and green space are associated with positive outcomes
like reduced smoking, alcohol consumption, obesity, or cognitive
decline (Jennings and Bamkole, 2019; Wendelboe-Nelson et al.,
2019).

Green spaces contribute to social cohesion through fostering


positive social interactions and social engagement (Jennings and
Bamkole, 2019). Natural features also enhance feelings of place
attachment and identity, promoting a sense of community that
contributes to a decrease in feelings of loneliness (Prezza et al.,
2001). A lower presence of green spaces in people's living
environment was found to be related to greater feelings of
loneliness and perceived shortage of social support (Maas et al.,
2009). The association between green spaces, perceived social
support and loneliness was found to be the strongest in highly
urbanized areas (Maas et al., 2009).

These research results, as well as the existing reality in the city


led the Connecting Nature team to consider Chronic loneliness as
a significant indicator to know the influence of the Bellahouston
Demonstration Garden (Figure 5-24) on the well-being of its users.

Description of The indicator is assessed using a standardized quantitative


Additional Indicator instrument: The Three-Item Loneliness Scale (Hughes et al.,
Application 2004). This tool is a short form of the revised UCLA Loneliness
scale (Russell et al., 1980) which measures the experience of
loneliness. This scale includes three items measured on a 3-point
Likert scale (1 = hardly ever; 2 = some of the time; 3 = often).
For final scoring purposes, each person’s scale responses to the
three items are summed, with higher scores indicating greater
experienced loneliness (Hughes et al., 2004).

Methodology and data analysis require high expertise in psycho-


social research but quantitative data collection requires no
expertise. During the Connecting Nature project, the data
gathering is conducted after the NBS implementation, but it allows
making comparisons between different areas of the city or
population groups (i.e., users vs no users). It is suggested to
conduct two data collection waves to assess the longitudinal
effects over time.

Stakeholders Connecting Nature; Glasgow City Council; Glasgow Community


involved Planning Partnership; Data collection experts

Barriers Although the officers leading the Food Growing Strategy were
encountered and aware that the Bellahouston Demonstration Garden provided
lessons learned social, environmental, health and economic benefits, they had
difficulties both in reflecting these advantages in official papers,

232
and in holding conversations with the community and funding
bodies (Hölscher et al., 2019).

Therefore, within the Connecting Nature project a suitable


business model was identified to scale up and replicate the project
to other areas of the city (van de Sijpe et al. 2019). In this way,
the Connecting Nature project provided the knowledge to develop
food growing business within the Food Growing Strategy of the
city council, conducting conversations with the community and
identifying possible funding routes.

Case study authors Adina Dumitru1 ([email protected]), David Tomé Lourido1,


Susana Pablo1
1
University of A Coruña, Spain

References Hölscher, K., Lodder, M., Collier, M., Frantzeskaki, N., Sillen, D., Notermans,
I., Allaert, K., Dumitru, A., Connop, S., Vandergert, P., McQuaid, S.,
Quartier, M., van de Sijpe, K., Vos, P., Dick, G., Kelly, S., Mowat, L.,
Sermpezi, R., Dziubala, A., Madajczyk, N., and Osipiuk, A., Deliverable
5: Nature-based Solutions Framework for frontrunner cities, 2019,
CONNECTING Nature, Grant Agreement number 730222.
Holt-Lunstad, J., Smith, T.B., and Layton, J.B., ‘Social relationships and
mortality risk: a meta-analytic review’, PLoS medicine, Vol. 7, No 7,
2010, pp. 1-20.
Holt-Lunstad, J., Smith, T.B., Baker, M., Harris, T., and Stephenson, D.,
‘Loneliness and social isolation as risk factors for mortality: a meta-
analytic review’, Perspectives on Psychological Science, Vol. 10, No 2,
2015, pp. 227-237.
Hughes, M.E., Waite, L.J., Hawkley, L.C., and Cacioppo, J.T., ‘A short scale
for measuring loneliness in large surveys: Results from two population-
based studies’, Research on Aging, Vol. 26, No 6, 2004, pp. 655-672.
Jennings, V. and Bamkole, O., ‘The relationship between social cohesion and
urban green space: An avenue for health promotion’, International
Journal of Environmental Research and Public Health, Vol. 16, No 3,
2019, pp. 452.
Maas, J., Van Dillen, S.M., Verheij, R.A., and Groenewegen, P.P., ‘Social
contacts as a possible mechanism behind the relation between green
space and health’, Health and Place, Vol. 15, No 2, 2009, pp. 586-595.
Pantell, M., Rehkopf, D., Jutte, D., Syme, S.L., Balmes, J., and Adler, N.,
‘Social isolation: a predictor of mortality comparable to traditional clinical
risk factors’, American Journal of Public Health, Vol. 103, No 11, 2013,
pp. 2056-2062.
Prezza, M., Amici, M., Roberti, T., and Tedeschi, G., ‘Sense of community
referred to the whole town: Its relations with neighboring, loneliness, life
satisfaction, and area of residence’, Journal of Community Psychology,
Vol. 29, No 1, 2001, pp. 29-52.
Russell, D., Peplau, L.A., and Cutrona, C.E., ‘The revised UCLA Loneliness
Scale: concurrent and discriminant validity evidence’, Journal of
Personality and Social Psychology, Vol. 39, No 3, 1980, pp. 472-480.
van de Sijpe, K., Vos, P., Dick, G., Mowat, L., Dziubala, A., Zwierzchowska,
I., Vandergert, P., Jelliman, S., Connop, S., Nash, C., and González, G.,
Deliverable 9: An interim report on progress towards initiation of city-
wide nature-based solutions exemplars, 2019, CONNECTING Nature,
Grant Agreement number 730222.
Wendelboe-Nelson, C., Kelly, S., Kennedy, M., and Cherrie, J. W., ‘A scoping
review mapping research on green space and associated mental health
benefits’, International Journal of Environmental Research and Public
Health, Vol. 16, No 12, 2019, 2081.
White, J.T. and Bunn, C., ‘Growing in Glasgow: Innovative practices and
emerging policy pathways for urban agriculture’, Land Use Policy, Vol.
68,2017, pp. 334-344.

233
Figure 5-24. Bellahouston Demonstration Garden (© Glasgow City Council).

5.3 Conclusions

The case studies herein illustrate the strength of the ‘buffet’ style approach of the
NBS impact indicator framework presented in this handbook. The inherent
heterogeneity of NBS – in type, form and scale of application – preclude a one-
size-fits-all approach to NBS impact assessment. In this context, the
Recommended indicators provide a suggested minimum suite of indicators in
order to obtain a holistic assessment of NBS performance and impact, with the
selection of specific Additional indicators serving to address specific concerns and
thus augment the achieved understanding. The preceding case studies show how
a combination of Recommended and Additional indicators may be applied to a
specific NBS in order to develop a comprehensive understanding of NBS
performance and impact, thereby enabling adaptive management of the NBS
asset.

234
NAIAD
Nature Insurance value: Assessment and Demonstration
Thames basin (GB) Medina del Campo aquifer (ES) Lower Danube basin (RO)
Lez basin (FR) La Brague basin (FR) Glinscica catchment (SI)
Copenhagen (DK) Lodz (PO) Rotterdam (NL)

NAIAD is aimed to develop a strong conceptual framework for evaluating the assurance and the
insurance value of ecosystem services. The project has developed the concept of natural assu-
rance schemes, and the range of tools and methods to design them. These range from physical,
social and economic assessments, integration and co-design with stakeholders, to the develop-
ment of business models and financing arrangements to their full implementation and monito-
ring. Stakeholders involved included insurers, river basin agencies, local authorities, farmers in
the validation and application in nine case study sites across Europe. It finally aims to contribute
to academic knowledge and policy action on NBS planning and integration, and contribute to
raise awareness on NBS and the associated socio-economic opportunities at all scales.

Drawing on knowlegde from projects


funded by the European Union
Image: Petit Buëch - Photo © Smigiba.fr
SCOPE NBS for Disaster Risk Reduction, i.e. floods and droughts

Approach to Impact Assessment Main Challenges addressed

The NAIAD framework is designed for effectiveness


assessment and decision-making with respect to
1. Climate Resilience
the choice of best NBS measures and strategies.
The different steps of disaster risk reduction and 2. Water Management
contributions of NBS are studied within the NAIAD
project considering technical, physical but also so- 3. Natural and Climate Hazards
cial, human, environmental and economic features.
A specific methodology is designed to determine 4. Green Space Management
the indicators. Relevant indicators are defined by
experts and stakeholders through workshops. A 5. Biodiversity
two-level approach is proposed making a differen-
ce between technical analysis and decision-making 6. Air Quality
contexts. Expert and technical assessments are
used as inputs in a multicriteria decision-making 7. Place Regeneration
framework which allows to address all kinds of tech-
nical, environmental, economic, or social features, 8. Knowledge and Social Capacity Building
and to consider stakeholder preferences as identi-
fied during participative workshops. 9. Participatory Planning and Governance

10. Social Justice and Social Cohesion


Involved Stakeholders and roles
11. Health and Wellbeing
A core operating principal of NAIAD is to proacti-
vely engage with stakeholders in the case studies
12. New Economic Opportunities & Green Jobs
throughout the application of its conceptual and
assessment methodologies for Natural Assurance
Schemes. The interdisciplinary nature of the whole
approach fundamentally makes it relevant to a wide Lessons learned
range of stakeholders, including decision makers,
practitioners, scientists, end users and communities. The first lesson learned on impact assessment
Each stakeholder will have their own particular know- from the NAIAD project is the importance of tai-
ledge and perspectives of the integrated physical, so- loring the approach to the catchment or pilot pe-
cial, cultural and economic systems in which the case culiarities. Providing an objective, easily unders-
study is situated, with all these needing to be shared tandable method to assess indicators of physical,
and synthesised during the assessments. In addi- social and economic effectiveness of NBS is es-
tion, the stakeholders served an important function sential to guarantee security but also to increase
in terms of “road testing” and validating the tools and acceptance by stakeholders.
methods developed and presented in this volume.
Different tools for impact assessment developed in
Municipal Administrations NAIAD are tailored to the different demos allowed
to get specific results for consensually agreed im-
pact indicators, with high level of acceptance and
Regional/national statistics authority satisfaction from stakeholders considering both
technical, physical, environmental, economic, soci-
Planning experts al and human effects and co-benefits of measures
and strategies. One example is the Flood-Ex-cess-
Scientists / Academia Volume (FEV) method that has been developed
to quickly assess cost-efficacy of flood-mitigation
NGOs strategies and proved useful in stakeholder work-
shops for raising public awareness of flood risk as-
River basin authorities sessment before choosing a NBS strategy.

Insurance sector
Learn more
Farmers www.naiad2020.eu

The NAIAD project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 730497
OPERANDUM
Open-air laboratories for Nature Based Solutions
to manage hydro-meteo risks
OAL-Australia OAL-Austria OAL-ChinaMainLand OAL-ChinaHongKong
OAL-Finland OAL-Germany OAL-Greece OAL-Ireland OAL-Italy OAL-UK

OPERANDUM will deliver tools and methods for the demonstration and market uptake of Natu-
re-Based Solutions to reduce hydro-meteorological risks. Nature-Based Solutions (NBS) are solu-
tions that are inspired and supported by nature. These solutions provide environmental, social and
economic benefits and help build resilience by bringing natural features into cities and landsca-
pes. In the OPERANDUM project, site-specific and innovative NBS are co-designed, co-developed,
deployed, tested and demonstrated with partners and local stakeholders in open-air laboratories.
These open-air laboratories (OALs) are natural and rural Living Labs that cover a wide range of
hazards with different climate projections, land use and socio-economic characteristics.

Drawing on knowlegde from projects


funded by the European Union
Image: Bellocchio Lagoon (OAL-Italy) - Photo © Filippo Zanni for Arpae, 2016
SCOPE To provide evidence, best practices, replication and scalability of existing
and novel NBS, foster uptake of solutions to increase market exploitation

Approach to Impact Assessment Main Challenges addressed

The project’s approach is based on 10 Open-Air La-


boratories: areas exposed to specific hydro-meteoro-
1. Climate Resilience
logical risks where the efficacy of existing and novel
NBS are assessed at local scale. OALs provide con- 2. Water Management
crete, flexible and transportable frameworks in order
to expand the adoption of green/blue/hybrid infras- 3. Natural and Climate Hazards
tructures across Europe and in developing countries.
The OALs in OPERANDUM demonstrate NBS for diffe- 4. Green Space Management
rent climatic zones and different climate change sce-
narios in Europe. The implemented NBS build upon 5. Biodiversity
multi-disciplinary expertise and full understanding of
the role of ICT and takes into account market ex- 6. Air Quality
ploitation and national, EU and international policies.
7. Place Regeneration
Involved Stakeholders and roles 8. Knowledge and Social Capacity Building

Due to the complexity of the Project a multiple level 9. Participatory Planning and Governance
structure of engagement strategy is required. Start-
ing from the local community, the Project involves 10. Social Justice and Social Cohesion
stakeholders at national and international level to
leverage widest possible NBS acceptance to promo- 11. Health and Wellbeing
te its diffusion as a good practice and push business
exploitation. The stakeholder engagement strategy
12. New Economic Opportunities & Green Jobs
is based on the stakeholder mapping to identify the
main target categories of OPERANDUM. An import-
ant step in the stakeholder engagement process is
represented by the prioritizing of stakeholders: a Lessons learned
Power-Interest Matrix has been adopted as a use-
ful tool to assessing the level of engagement requi- The challenges found across the OALs so far (OPER-
red of different stakeholder groups. Furthermore, NADUM is still halfway) are related to the awareness,
for each stakeholders category, reasons of interest attitudes and trust, diversity of goals and interests,
and expectations have been identified to obtain a financial, legislative, resources (skills or time). We
greater understanding of stakeholders motivations, found that monitoring during the co-creation process
interests, needs, and requirements. and evaluation at the end of the process are im-
portant phases to faciltate the adoption of changes,
improve the process, and enhance learning among
Municipal Administrations partners. Defining common strategy for stakeholder
engagement that includes tactics, formats, ethical
Citizen rules and indicators for monitoring, is found to be
of a paramount importance. The involvement of sta-
keholders has to be promoted in every step of the
Planning experts project and it’s essential to maintain current commu-
nication or collaboration practices according to the
Scientists / Academia needs of each phase. The novel platform, the OPE-
RANDUM-GeoIKP has been designed ad-hoc to reach
Green businesses target users (stakeholders) including citizens, public
authorities, policy makers. It is mandatory that in-
Regional/national authority formation is conveyed using the up-to-date scientific
evidence as well as worked examples.
National/regional park‘s authorities

International bodies
Learn more
Policy makers www.operandum-project.eu

The OPERANDUM project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 776848.
PHUSICOS
Solutions to reduce risk in mountain landscapes
Gudbrandsdalen Valley (NO) The Pyrenees (ES/FR) Isar River Basin, Munich (DE)
Serchio River Basin / Massacciuccoli Lake (IT) Kaunertal Valley (AT)

PHUSICOS, meaning ‘According to nature’, in Greek φυσικός, aims to demonstrate how natu-
re-inspired solutions reduce the risk of extreme weather events in rural mountain landsca-
pes. The focus of PHUSICOS is on demonstrating the effectiveness of NBS and their abili-
ty to reduce the impacts from hydro-meteorological hazards (flooding, landslide, erosion,
drought, snow avalanche) in rural mountain landscapes. The NBS considered and imple-
mented in PHSUICOS are cost-effective and sustainable measures inspired by nature that
attenuate, and in some cases prevent, the impacts of natural hazard events and thereby the
risks that affect the exposed regions.

Drawing on knowlegde from projects


funded by the European Union
Image: Small landslide in Kvam, Norway - Photo © Ingar Haug Steinholt, NGI
SCOPE NBS for disaster risk reduction in rural mountain landscapes

Approach to Impact Assessment Main Challenges addressed

The PHUSICOS NBS Impact Assessment Frame-


work is based on a multicriteria decision analy-
1. Climate Resilience
sis, which assesses, through a matrix containing
indicators aggregated in different sub-criteria, 2. Water Management
the risk reduction performance and the co-be-
nefits of a design scenario for a specific site. In- 3. Natural and Climate Hazards
dicators are selected after an extensive review
of the main existing NBS project networks and 4. Green Space Management
platforms, as well as the challenges indicated by
the EKLIPSE project. The five main categories 5. Biodiversity
(ambits) considered in the evaluation of an NBS
in the PHUSICOS framework are 1) Risk reduc- 6. Air Quality
tion, 2) Technical and feasibility aspects, 3) En-
vironment and ecosystems, 4) Society, and 5) 7. Place Regeneration
Local Economy.
8. Knowledge and Social Capacity Building

9. Participatory Planning and Governance


Involved Stakeholders and roles
10. Social Justice and Social Cohesion
Stakeholder involvement and participation is a
key component in the successful design, planning 11. Health and Wellbeing
and implementation of NBS. PHUSICOS uses a Li-
ving Labs approach to frame and carry out the 12. New Economic Opportunities & Green Jobs
participatory processes with stakeholders at the
different case study sites. Rather than a single de-
finition, PHUSICOS has emphasized focusing on
Living Lab principles to ensure tailor-made pro-
Lessons learned
cesses for co-creating and co-developing NBSs
including fostering innovation and learning, diver- As part of the process of monitoring relevant
sity, user-centered, locally relevant context, and indicators to assess the impact and efficacy of
open-mindedness. The PHUSICOS Living Labs also NBSs, stakeholders in the Living Labs have been
highlight the need to engage stakeholders from engaged to provide input to the development of
four main networks: public organizations, private these monitoring systems. Thus far, reflections
companies, users (or end-users), and knowledge have been collected from the Serchio River Basin
institutions (academia). These different groups of demonstrator case study site at Massacciuccoli
stakeholders are providing initial reflections and Lake in Italy. In dialogue with local farmers, buffer
identifying indicators that are most relevant ba- strips to reduce the hydro-meteorological risk and
sed on their knowledge and needs with regard to improve the water quality are being implemen-
implementing and monitoring NBS. ted. Feedback on monitoring indicates that for
each of the five main categories (ambits) in the
PHUSICOS NBS evaluation framework, at least
one of the proposed indicators is considered use-
Municipal Administrations ful; with those focusing on implementation and
maintenance costs as well as the policy context as
Regional/national statistics authorities the most valuable. Furthermore, publicly sharing
monitoring results is viewed positively, also as a
Citizen means of promoting NBS to the public.

Planning experts

Scientists / Academia
Learn more
www.phusicos.eu

The PHUSICOS project has received funding from the European Union‘s Horizon 2020
research and innovation programme under grant agreement No 776681.
RECONECT
Regenerating Ecosystems with Nature-based
solutions for hydro-meteorological risk rEduCTion
Elbe Estuary (DE) Seden Strand Odense (DK) Todera River Basin (DK) Park Portofino (IT)

Ijssel River Basin (NL) Inn River Basin (AT) Greater Aarhus (DK) Thur River Basin (CH)

Var River Basin (FR) Les Boucholeurs (FR) Kamchia River Basin (BG) Pilica River Basin (PL)

Sava River Basin (RS/HR) Chao Praya River Basin (TH) Greater Tainan Coastline (TW)

Rio do Couves (BR) Klang River Basin (MY) Yangtze River Basin (CN) Chindwin River Basin (MM)

Tarago River Basin (AU) Trinity River Basin (US) Piura River Basin (PE) Rio Frio (CO)

Cañaveralejo, Lili and Melendez River Basins (CO) Coastline of St. Maarten (SX)

RECONECT aims to rapidly enhance the European reference framework on Nature-Based Solu-
tions (NBS) for hydro-meteorological risk reduction by demonstrating, referencing, upscaling and
exploiting large-scale NBS in rural and natural areas. In an era of Europe’s natural capital being
under increased cumulative pressure, RECONECT will stimulate a new culture of co-creation of
‘land use planning’ that links the reduction of hydro-meteorological risk with local and regional
development objectives in a sustainable and financially viable way.To do that, RECONECT draws
upon a network of carefully selected Demonstrators and Collaborators that cover a wide and di-
verse range of local conditions, geographic characteristics, institutional/governance structures
and social/cultural settings to successfully upscale NBS throughout Europe and Internationally.

Drawing on knowlegde from projects


funded by the European Union
Image: Seden Strand Odense - Photo © RECONECT Project
SCOPE RECONECT demonstrates, references and
upscales Nature-Based Solutions in rural and natural areas

Approach to Impact Assessment Main Challenges addressed

In RECONECT, NBS Impact Assessment is carried


1. Climate Resilience
out in relation to three categories of challenges
i.e., WATER, NATURE and PEOPLE. Where possib-
le, monitoring data is being, or will be, collected 2. Water Management
and transmitted through real-time SCADA/tele-
metry services and also through social science 3. Natural and Climate Hazards
surveys. These data will be used to evaluate the
NBS impacts in relation to benefits, co-benefits as 4. Green Space Management
well as the negative effects.
5. Biodiversity
Monitoring and evaluation of NBS against the
WATER challenges address questions related to 6. Air Quality
hydro-meteorological risks. Monitoring and eva-
luation of NBS against the NATURE challenges ad- 7. Place Regeneration
dress questions related to habitat structure and
the biodiversity of flora and fauna. Monitoring and 8. Knowledge and Social Capacity Building
evaluation of NBS against the PEOPLE challenge
address questions concerning social and econo- 9. Participatory Planning and Governance
mic benefits, with implications for human health
and well-being and resilience to impacts from hy- 10. Social Justice and Social Cohesion
dro-meteorological events.
11. Health and Wellbeing

12. New Economic Opportunities & Green Jobs


Involved Stakeholders and roles

A co-monitoring and co-evaluation framework is


being developed for Demonstrators A and B. There Lessons learned
are two kinds of RECONECT monitoring activities
within this framework. The first one is monitoring There is some information available that can
to assess the state of the system (e.g. the general be used to evaluate the impact of NBS on hy-
conditions in the NBS area), i.e., baseline monito- dro-meteorological risk reduction and biodiver-
ring before construction of NBS, and the second sity enhancement. However, there is still a lack
one is monitoring to assess the performance of of knowledge in terms of monitoring and impact
implemented NBS towards the achievement of the evaluation for PEOPLE benefits (e.g., human
project’s goals/sub-goals. health and well-being).

Municipal Administrations (FR/FL)

Regional/national statistics authority

Citizen

Planning experts

Scientists / Academia
Learn more
www.reconect.eu
NGOs

The RECONECT project has received funding from the European Union‘s Horizon
2020 research and innovation programme under grant agreement No 776866
6 NBS FOR DISASTER RISK REDUCTION
Coordinating Lead authors
Nadim, F., Tacnet, J.-M.

Contributing authors
Basco Carrera, L., Capobianco, V., Caroppi, G., Gerundo, C., Giugni, M., Manojlovic, N., Oen,
A., Pilla, F., Piton, G., Porcu, F., van Cauwenbergh, N., Scheuer, S., Vojinovic, Z.

Summary

What is this chapter about?

Losses and damages due to natural hazards can be dramatic. This chapter
provides a global overview of the requirements for risk assessment in the context
of Disaster Risk Reduction (DRR). It outlines how NBS as structural measures can
effectively reduce risks related to hydro-meteorological disasters, at the same
time providing multiple co-benefits. As NBS may lack sufficient physical capacity
to provide adequate protection against extreme events, the chapter illustrates
how in most cases a hybrid combination of NBS and technical engineering (i.e.,
green and grey) measures can provide the optimal solution when DRR is the
primary goal.

Next, we introduce the assessment of effects and co-benefits of NBS. These co-
benefits should be included in cost-benefit analyses when comparing NBS with
grey or hybrid solutions. Case studies illustrate selected implementation
pathways and exemplify indicators and assessment frameworks that can be used

243
to assess different aspects of technical, physical, economic, social, human and
environmental features of NBS.

How do I use this chapter in my work with NBS?

The frameworks, indicators and case study examples provided in this chapter can
be used to design a monitoring and evaluation system for an existing or planned
NBS for DRR.

When can I use this knowledge in my work with NBS?

Assessing the effectiveness of NBS at regional or local level for DRR in the context
of hydro-meteorological hazards requires a detailed assessment of the risk level
and the expected impact of the implementation of NBS. The knowledge presented
in this chapter will assist in designing the monitoring and evaluation system for
this purpose, including the selection of appropriate criteria and methods.

How does this chapter link with the other parts of the handbook?

This chapter expands the discussion of NBS impact evaluation from the city scale
(chapters 1-5) to the catchment scale in the context of large-scale NBS for
disaster risk reduction, with a primary focus on hydro-meteorological risk
reduction.

6.1 NBS and Disaster Risk Reduction

As mentioned in the opening sentence of Chapter 1, urban areas cover less than
4% of land all around the world. Yet, almost all of the NBS-related research
projects funded by the European Commission (EC) before 2018 focused on
problems in urban areas. Nearly 50% of the rural areas in the world are classified
as mountainous regions and are exposed to risk from geological and hydro-
meteorological hazards. Mountains tend to amplify these risks, and even more so
under extreme weather events. However, rural mountainous regions do not
receive the same attention as densely populated urban areas in national disaster
risk reduction (DRR) plans. National DRR plans focus mainly on regions with
highest population density, which tend to be urban and/or coastal areas. Impacts
of extreme hydro-meteorological events in mountain areas often affect entire
river basins. Some of the natural hazard-related disasters in urban and coastal
areas such as flooding caused by landslide dam breaks during and after storms
are due to processes and events like flash floods and landslides that initiate in
hilly and mountainous regions higher up in the river basin. Nature-based
Solutions (NBS) have many advantages to fulfil disaster risk reduction objectives
but their implementation is still limited because of lack of evidence of their
effectiveness. Four recent H2020 projects – NAIAD, PHUSICOS, OPERANDUM and
RECONECT – focus fully or partially on demonstrating the effectiveness of nature-
based solutions and their ability to reduce the impacts from small, frequent
events (extensive risks) in rural mountain landscapes and in coastal areas. To
demonstrate the effectiveness of NBS in achieving DRR objectives and to measure

244
their co-benefits, specific methodologies and measurable indicators are needed
to provide evidence to stakeholders and decision-makers.

The previous chapters of this handbook review the existing indicators for all
environmental challenges in which NBS may be considered. However, it appears
that the existing frameworks related to indicators for measuring the effectiveness
NBS only partially address the issue of disaster risk reduction. Evaluating the
effectiveness of risk reduction measures, and especially NBS, requires
understanding and describing the effects of measures (i.e., their physical
capacity) on phenomenon’s nature, intensity and frequency. The concept of
effectiveness itself and the related indicators are linked to the comparison of an
objective assigned to a function and a capacity (see Chapter 2).

In the critical domain of disaster risk reduction, demonstrating the physical


effects of those measures is therefore a first essential step towards their
successful implementation. However, in addition to this somewhere classical and
expected effect, NBS can offer other co-benefits that conventional grey
infrastructures (e.g., dams, levees) do not provide in terms of environmental,
economic, social co-benefits. Indicators in the DRR context are not only physical;
they should include other categories like risk perception, environmental impacts,
and economic effectiveness.

This chapter extends the existing framework and proposes to address this
challenging topic by taking benefit of recent projects dedicated to hydro-
meteorological risks, mainly NAIAD, PHUSICOS, RECONECT and OPERANDUM 51.
It first recalls briefly natural risks contexts, basics of risk assessment, risk
reduction measures and then describes relevant indicators and principles that
should guide indicator selection for disaster risk reduction. It focuses on the role
of NBS for adaptation to and mitigation of impacts of weather events – with some
examples taken from projects’ representative case studies.

6.2 Basics of risk analysis, risk reduction measures, resilience and


effectiveness

Defining, selecting and assessing indicators of NBS effectiveness in the context


of DRR is linked to the understanding of risk concepts and the possible effects
that NBS may have on those risk components. Depending on phenomena (e.g.,
floods, mountain-flash floods, debris-flows, landslides, rock falls), the physics of
hydrological, geophysical processes may differ, although a common approach can
be applied. This section presents those common points.

Risks result from a combination of hazard (frequency and intensity), exposure


and potential losses as a function of vulnerability and values. Here, vulnerability
represents the degree of damage or loss when an exposed element such as an

51
NAIAD: http://naiad2020.eu/; PHUSICOS: https://phusicos.eu/; RECONECT: https://reconnect-
europe.eu/; OPERANDUM: https://reconnect-europe.eu/

245
object, a person, or an activity is impacted by a given level of phenomenon
intensity (Figure 6-1).

Intensity depends on the considered phenomenon and its several possible effects.
For instance, mountain floods are not only composed of water but also transport
solids (sediment and large wood). Measuring only water height may, therefore,
not be relevant for computing damages, first, because of bed level change due
to deposition or erosion, and, second, because these changes and/or damages
due to material load may be the main cause of damage rather than the mere
submersion by water (Figure 6-2).

Figure 6-1. Basic components of risk: the effectiveness of a risk reduction measure requires to analyse its
effects on the phenomenon including (1) the nature of the effects (e.g., flooding, scouring, impact of
boulders); (2) their frequency; and (3) their intensity (e.g., flood depth) and their interaction with exposed
elements (exposure and vulnerability as a potential of damage).

246
Figure 6-2. Positive and negative effects of NBS on phenomena and protections’ physical features are
addressed to assess measures’ effectiveness (Tacnet, 2019).

Risk reduction measures consist of both structural (physical) measures such as


protective structures (e.g., check dams) or non-structural measures such land-
use management, land-cover control and risk mapping (Figure 6-3). Structural
measures aim to reduce risk by having a physical effect on the main
characteristics of a phenomenon (e.g., reducing run-off on a given territory).

Figure 6-3. When dealing with DRR, Nature-Based Solutions are part of structural risk reduction strategies.

247
Nature-based solutions can therefore be considered as a structural measure
dedicated to having an effect on the hazard component of risk (i.e., on the
frequency or intensity of a given phenomenon). According to Evette et al. (2009),
living plants have been used for a very long time throughout the world in
structures against soil erosion, as traces have been found dating back to the first
century BC. In Western Europe, bioengineering was widely practiced during the
eighteenth and nineteenth centuries. For instance, since the 19th century in
France, soil restoration, protection forests, gully restoration and planting as well
as torrent check dams have been aiming to reduce sediment production and risks
to people and assets in the valleys. Many techniques and hybrid combinations
with civil engineering solutions are therefore not new (Figure 6-4). However,
characterising the effectiveness of those measures remains difficult.

13.

Figure 6-4. Combination of civil-engineered solutions and reforestation (which can be defined as Nature-
Based Solutions) have been experimented successfully since the 19th and 20th century for mountain
restoration purpose 52, here with an example from the south eastern French Alps.

52
See Restaurer la montagne. Photographies des Eaux et Forêt du XIXe siècle. Brugnot, G., Coutancier, B. et
al., Paris: Somogy éd. d'art, ISBN: 2-85056-801-5, 188 p.

248
For flood risk management, many types of NBS exist, each of them corresponding
to a specific expected function that will be analysed to check their effectiveness 53
through their comparison between their physical capacity (e.g., a storage
volume) and an objective linked to this function (e.g., volume needed) (Figure
6-5).

14.

Figure 6-5. NBS used for flood risk management have different functions.

53
See chapter 2 for a definition of effectiveness

249
6.3 Indicators and methodologies for measuring NBS effectiveness
indicators in DRR context

Several recent H2020 projects address the analysis of the effects of NBS. NAIAD,
PHUSICOS, RECONECT and OPERANDUM projects propose generic assessment
frameworks for measuring the effectiveness of an NBS that is primarily designed
for DRR.

The NAIAD framework is designed for effectiveness assessment and decision-


making with respect to the choice of best NBS measures and strategies. The
different steps of disaster risk reduction and contributions of NBS are studied
within the NAIAD project considering technical, physical but also social, human,
environmental and economic features (Figure 6-6). A specific methodology is
designed to determine the indicators. Relevant indicators are defined by experts
and stakeholders through workshops. A two-level approach is proposed making
a difference between technical analysis and decision-making contexts. Expert and
technical assessments are used as inputs in a multicriteria decision-making
framework which allows to address all kinds of technical, environmental,
economic, or social features, and to consider stakeholder preferences as
identified during participative workshops (Figure 6-7).

Figure 6-6. NAIAD’s global framework to assess role of NBS in Disaster Risk Reduction (DRR).

250
Figure 6-7. A multicriteria decision-making framework allows to integrate and combine technical, physical,
environmental and economic indicators. Decision makers express their preferences on high-level criteria
(protection level, economy of projects, social/cultural and environmental impacts). Experts provide and
assess indicators for those categories (adapted from Tacnet et al., 2018).

Regarding DRR, the indicators for measuring the NBS effectiveness in the NAIAD
framework are linked to physical effects of measures at different scales. The
NAIAD framework is applied ex-ante, the indicators related to physical effects are
thus assessed by a combination of numerical modelling and geomorphological
analysis. NAIAD proposes a global hierarchical model to combine indicators for
various aspects including technical, physical, organisational, environmental,
social/human and economic features (so-called TOPHEE approach) in order to
assist the decision-making process.

The projects PHUSICOS, RECONECT and OPERANDUM all focus on NBS for
reducing the risk of hydro-meteorological hazards. However, they approach the
problem from different viewpoints and their recommended frameworks have their
distinct characteristics. Table 6-1 compares some of the characteristics of these
frameworks. All three frameworks are built on the basis of the hazards addressed
in the case study sites of each project. For example, RECONECT focuses only on
flood and drought risk; PHUSICOS on landslides, snow avalanches, floods and
drought; and OPERANDUM focuses on a larger spectrum of hazards (Shah et al.,
2020), including coastal erosion, storm surge, nutrient and sediment
accumulation, soil salinization, heat waves, and dust storms found in the Open
Air Laboratories (OAL).

251
PHUSICOS and RECONECT have both selected the risk and co-benefits categories,
as well as the initial set of indicators to be assessed on the basis of existing NBS
projects, platforms and literature, with a focus on the challenges indicated by the
EKLIPSE project. A different approach was adopted by the OPERANDUM team,
who identified the indicators through the review of literature available for each of
the OAL-specific hazards, together with stakeholder involvement in surveys and
focus group discussions. In the OPERANDUM framework, once the potential
indicators are identified, their final selection is based on four criteria: Credibility,
Salience, Legitimacy, and Feasibility. Stakeholders are involved in all processes,
from the co-design of the framework to the co-selection of the indicators, based
on their specific needs and priorities. The OPERANDUM framework has not been
tested yet, while the other two frameworks were tested on a real NBS case in
Thailand within the RECONECT project, and for three hypothetical scenarios in
PHUSICOS: (1) the Baseline Scenario before implementation of any mitigation
measure; (2) a NBS Scenario; and (3) a Hybrid Scenario.

Based on the tests carried out to date, it can be noted that RECONECT approach
has been solely used for ex-post assessment of a NBS scenario for potential
replication, up-scaling or improvement. This is different from the PHUSICOS
framework, which can be used also as a decision-making tool to compare the
potential performances and co-benefits of different design scenarios for a specific
context prior to their implementation. A main feature of the RECONECT
framework is that each indicator is expressed in a relative manner, i.e., as the
difference between its value in the NBS scenario and in the scenario without NBS,
whilst for PHUSICOS and OPERANDUM the indicators are expressed using
absolute values. This difference highlights the importance that the RECONECT
project attributes to the NBS co-benefits. PHUSICOS and OPERANDUM
systematically address the risk reduction provided to a specific context, in terms
of changes in exposure, vulnerability and hazard. Furthermore, the OPERANDUM
framework treats both the ecosystem and the society as elements exposed to
risks posed by hydro-meteorological hazards at each specific OAL, highlighting
again the adopted risk-oriented approach.

A general observation that can be made is that the RECONECT framework is


benefits-oriented, the OPERANDUM framework is risk-oriented, and the added
value of the PHUSICOS framework is balanced and neutral: risk reduction
indicators and co-benefits indicators are structured in a way that the stakeholders
can state their preferences through weights assigned to each indicator. An added
value of the OPERANDUM framework is that it is applicable from a local to
regional/national scale, while both RECONECT and PHUSICOS are mostly focused
at local or catchment scale. An added value of the RECONECT framework is that
it includes, as last step of the evaluation, an analysis of a so-called NBS grade,
focusing on the weakest indicators, so that experts and stakeholders can provide
recommendations for all indicators, or only those with low scores.
Recommendations can include guidance on how to better involve stakeholders in
every step of the framework, how to better measure, collect, and analyse data,
and how to maintain the NBS to maximize benefits. Finally, all three frameworks
are highly flexible, and they can be adapted or redefined to the context where
they are applied, depending on the needs of the stakeholders and the most
suitable indicators to be assessed.

252
Table 6-1. Key features of the frameworks developed in EC H2020 hydro-meteorological risk reduction projects (based on partial examples presented in case studies).

Framework aspect NAIAD RECONECT PHUSICOS OPERANDUM

Key features of the Integrated hybrid Five main sequential Based on a multicriteria Vulnerability and risk
frameworks approach mixing classical steps, from the selection decision analysis (MCDA), assessment framework,
engineering, and the evaluation, to the which assesses, through aimed at looking at the
environmental and scoring of the main a matrix containing impacts of hydro-
geomorphological indicators for the indicators aggregated in meteorological hazards
approaches but also assessment of the different sub-criteria, the on an exposed social-
systemic analytical, benefits of an risk reduction ecological system
economic and implemented NBS performance and the co-
multicriteria decision- benefits of a design
aiding frameworks scenario for a specific site

Source for the Multidisciplinary Indicators as well as the Indicators are selected Systematic literature
identification of the indicators, either from three benefit categories after an extensive review review combined with
initial set of indicators existing methods (e.g., where they fall in (Water, of the main existing NBS stakeholders and expert
EU Reform project for Nature, People), based on project networks and surveys and focus group
morphological quality the challenges indicated platforms, as well as the discussions
index) or self-created by the EKLIPSE project challenges indicated by
(e.g., flood excess the EKLIPSE project
volume, FEV)

Type of hazards Flood Flood, drought Flood, landslide, snow Hydro-meteorological but
addressed avalanche, drought can be applied to any
natural hazard

Main categories Integrated risk Water, Nature, People Risk reduction, Technical All components of risk
management, and feasibility aspects, (hazard, exposure (social
Multifactorial NBS Environment and and ecological sub-
effectiveness assessment, ecosystems, Society, systems), and
Decision-aiding Local Economy vulnerability (social
sensitivity and coping
capacity; ecosystem

253
sensitivity and
robustness)

Indicator types Multicriteria technical, Relative value Absolute value Not specified, but
physical, environmental, absolute value is implied
social/human, followed by normalization
organisational indicators
(relative, comparative
and absolute values) –
TOPHEE approach

Stage of assessment Ex-ante assessment Ex-post assessment (can Ex-ante assessment and Ex-ante (can be
also be applied for Ex- Ex-post assessment visualised, e.g., with
ante assessment) scenario development)

Spatial scale of Local or catchment scale Local or catchment scale Local or catchment scale Local to basin scale
application (can be extended to
regional or global scale)

Environmental context Urban and rural environmental contexts

Stakeholder level of Stakeholders are involved Stakeholders are involved Stakeholders are involved Stakeholders are
involvement in the indicator selection in the process from step 1 in the refinement of the continuously involved,
process (workshop), the (selection of indicators) to matrix for the specific they help to co-design
assessment process step 4 (evaluation of the site, as well as in the framework, co-select
(validation, NBS grade). It is not weighing the ambits, the indicators, and give a
communication of specified if they are criteria and indicators prioritized list of
technical assessment), actively involved also in indicators. They will be
the decision-aiding step step 5 involved in weighing
(identification of (recommendations) indicators
preferences, assessment
of solutions)

254
Outcome Fully integrative and NBS grade incorporating Overall scenario scoring Risk to the social-
versatile framework from all the benefits assessed, for comparing two ecological system
indicators design to their equal to the average of different scenarios, or to
aggregation, NBS the scores of each assess a specific scenario
strategies and measures indicator quantified performance over time
are assessed in a
multicriteria perspective

255
6.4 Case study #1 - NAIAD (La Brague, FR): from indicators
assessment to integration and decision-aiding for flood risk
management 54

6.4.1 Context and global framework for assessment of NBS effectiveness

Several scales and kinds of application test cases were considered in the NAIAD
project 55. This case focuses on La Brague River in the south of France, where the
effectiveness of nature-based solutions was addressed through a combination of
physical, geomorphological and economic indicators. The Brague River basin is a
68 km² catchment located along the French Mediterranean coast between the
cities of Cannes and Nice. The Brague is a short river, 21 km long, and is
subjected to flash floods as well as woody debris production and transport.
Mediterranean climate causes heavy rains mostly in autumn, and the floods of
the Brague are often devastating and sometimes deadly. Over the period of
1970–2015, the Brague caused fourteen disastrous floods and eight deaths. The
insured damages of the October 2015 flood (which had an estimated return
period of over 100 years) amount to about 50 million € in the municipalities of
Biot and Antibes. After this flood, several campsites located in the area were
closed by state decision due to risk of being flooded. However, dozens of houses
remain at risk. This regrettable event provided an opportunity to re-define the
economic development strategy of the valley and to design new flood protection
strategies to both protect people and infrastructure against flood risk, and to
improve the river corridor’s natural life, landscape and environmental quality.

Risk analysis is traditionally addressed through hazard and vulnerability


assessment. The primary expectations of the selected nature-based strategies
for the river corridor would be that these strategies are effective in reducing
hazards from a physical point of view by storing water in the upper catchment
while easing drainage without overflowing in the lowlands. NBS can provide other
important co-benefits but they may appear as secondary if the protection level is
not sufficient. When used alone, eco-engineering approaches can propose
aesthetic solutions which may not be able to cope with required hydraulic capacity
or be strong enough to resist to hydraulic constraints. NBS flood alleviation
strategies studied for the Brague catchment are a combination of retention
measures by small natural retention areas in the upper catchment, along with a
widening of the river corridor in the lowlands enhanced by floodplain
reconnection. Floodplain works consist of several measures including bed and
bridge widening, forest corridor and wetlands restoration, and large woody debris
management. They are integrated in a so-called “giving-room-to-the-river”
strategy. Two levels of ambition, namely high and very high, are considered as
well as a more classical grey scenario based on huge retention dams for
comparison purposes.

NAIAD proposes both indicators and an original approach to formalize the concept
of effectiveness to design, assess and combine ad-hoc effectiveness indicators
(see also systemic analysis 56). A multidisciplinary approach draws on the

54
J.-M. Tacnet, G. Piton (INRAE/NAIAD)
55
See Deliverable 6.4 for an extended description of outputs
56
See NBS handbook, Chapter 2

256
knowledge of experts in forest and river management, natural hazards (floods,
erosion, wildfires), vulnerability and damage assessment, economy and decision-
aiding to perform an in-depth study of the Brague River catchment and compare
the effectiveness of possible grey (civil-engineered), green (nature-based
solutions) and hybrid strategies. Experts’ analysis and domain-specific methods
are used as basic inputs to address technical, environmental and economic
indicators. For instance, cost-benefit analysis is used to provide an indicator for
economic effectiveness assessment, morphological quality index (MQI; Rinaldi et
al., 2013) is used to assess the morphological status of the river while the flood
excess volume (Bokhove et al., 2019, 2020) is used to measure the physical
hydraulic capacity of measures and comparison with their economic features.
Total costs of the three protection strategies were evaluated and compared with
mean annual avoided losses (costs) based on historical events and theoretical
floods with known return period. The co-benefits related to NBS strategies were
also evaluated using two different methods (Arfaoui and Gnonlonfin, 2020a,
2020b). First, transfer of values based on a meta-regression-analysis of values
provided in other catchments, and second, a contingent valuation performed
locally through interviewing more than 400 persons in the basin. It should be
stressed that several intangible criteria, e.g., the improvement of the natural
status of the river, are poorly captured by the monetary methods and a
complementary multicriteria decision framework was developed to handle both
tangible and intangible criteria (Figure 6-7).

6.4.2 Indicators for assessment of technical, physical and economic efficacy


of flood mitigation strategies including NBS

Indicators to describe the environmental, ecological and geomorphological


status of a river have been estimated for the initial state and with assumptions
corresponding to the different NBS strategies and scenarios (Figure 6-8 and Table
6-2). The morphological quality index (MQI 57) aggregates 28 indicators
corresponding to geomorphological functionality, artificiality and channel
adjustments. It captures degradation of the geomorphological quality of the river
for the grey scenario (decreases in MQI, red cells in Table 6-2), whilst the NBS
scenarios improve it (increases in MQI, blue cells in Table 6-2). These elements
were meaningful for stakeholders interested in river restoration.

57
Developed within the EU project REFORM (https://reformrivers.eu)

257
Figure 6-8. Map of Morphological Quality Index values – state based on data and maps in 2017.

Table 6-2. Morphological Quality Index values for the different reaches (status 2017) and values of
intermediate aggregation. Those indicators are then used in the multicriteria decision-aiding framework.

Strategy Reach Brague Brague Brague Valmasque Valmasque


Gorges Biot #2 Antibes Gorges #5 Biot #4
#1 #6

Current MQIcurrent 0.94 0.82 0.80 0.94 0.85


status

Grey MQI 0.87 0.79 0.76 0.87 0.82

MQI-MQIcurrent=ΔMQI -0.07 -0.03 -0.04 -0.07 -0.03

NBS MQI 0.97 0.86 0.90 0.97 0.88


ambitious

MQI-MQIcurrent=ΔMQI 0.03 0.04 0.1 0.03 0.03

NBS very MQI 0.97 0.87 0.95 0.97 0.88


ambitious

MQI-MQIcurrent=ΔMQI 0.03 0.05 0.15 0.03 0.03


Note: MQI = 0 = river status totally altered; MQI = 1 = No alteration to the natural status

258
Providing an objective, easily understandable method to assess indicators of
physical and economic effectiveness of NBS is essential to guarantee security
but also to increase acceptance by stakeholders. The Flood-Excess-Volume
(FEV) method has been developed to quickly assess cost-efficacy of flood-
mitigation strategies by allowing generic flood-mitigation strategies to be tailored
to specific river-catchment scenarios. Produced through a collaboration between
the University of Leeds, UK, and the NAIAD project, it has been successfully
tested on data accrued from real flood events occurring in the UK (Aire and Calder
Rivers), France (La Brague, NAIAD demonstration) and Slovenia (Glinsisca river,
NAIAD demonstration), see Bokhove et al. (2019, 2020) and Pengal et al. (2020).

FEV identifies and utilises indicators of flood severity that are quantifiable, easy
to understand and to measure, hence making it objective, transparent to scrutiny
and user-friendly. It is repeatable, flexible and capable of rapidly verifying
whether or not a given ensemble of protection measures is sufficient to mitigate
against a priori specified degree of flood severity. The input data required by the
tool are the project-flood hydrograph (i.e., the water-discharge time series), the
water stage-discharge curve (i.e., the channel capacity) and the threshold level
(i.e., the discharge above which severe flooding occurs). In Figure 6-9, the
computed FEV represents the amount of water that cannot be contained by
existing flood defences for a given flood (Figure 6-9(1)). It then computes the
size of a virtual lake, 2 m deep and square in shape, that could retain the
computed FEV (Figure 6-9(2)). The last step is to split the lake into constituent
components, each of which is associated with a specific flood-protection measure
such as restored wetlands, leaky dams, floodplain reconnection, flood-retention
dams and giving-room-to-the-river, and to compare with their relative costs
(Figure 6-9(3)).

The tool has already proved useful in stakeholder workshops for raising public
awareness of flood risk assessment. This visualisation — of a virtual square lake
of human-scale depth — helps stakeholders to assimilate in a meaningful way the
excess of water that must be contained and/or confined in order to offer flood
protection. The simplified visualisation deliberately allows, and hence empowers,
a wide, non-expert audience to comprehend the magnitude of the amount of
water that needs to be contained/confined to mitigate flooding. The feedback
from end-users has been unanimous: the tool has unequivocally bridged the gap
between the design of local measures that were formerly unable to establish the
full picture of the catchment size flooding with advanced numerical modelling that
was, although powerful and precise, either too slow or too computationally
expensive to explore a plethora of potential protection strategies in a cost-
efficient manner.

In essence, it is by tailoring our approach to the catchment peculiarities using


relevant tools with various degrees of complexity that NAIAD helped decision-
making in the Brague catchment.

259
Figure 6-9. Different steps and results of Flood Excess Volume methodology: physical and economic
effectiveness of Nature-Based solutions are assessed and compared for different strategies.

6.5 Case study #2: A green barrier to reduce the risk of floods due
to snowmelt and extreme rainfall, Gudbrandsdalen Valley,
Norway

6.5.1 General background and hazard type

The Gudbrandsdalen is one of the most populated valleys in Norway. The valley
encompasses an area of ca. 15 km2 and is rich in floodplains along the river,
which are extensively used as farmland. Due to lack of other available land,
many settlements are located along the river. Historically, the valley is
susceptible to snowmelt flooding. However, this has been changing in recent
years with an increased risk of flooding due to heavy rainfall, also in
combination with snowmelt. Two major flood events in 2011 and 2013, causing
massive damages to infrastructure along the river (Figure 6-10), were the
driving factors behind the initiative to develop a Regional Master Plan for the
Gudbrandsdalen and its tributaries. The master plan proposes providing more
"room for the river" in flood-prone locations.

260
Figure 6-10. Valley of Gudbransdalen during the flood of 2013.

6.5.2 Co-benefits of the proposed NBS

The receded green barrier will provide space for the river during periods of
flooding, foster the natural processes in the watercourse and thus contribute
positively to the floodplain ecosystem. The landscape architect company
AgenceTer (PHUSICOS partner) highlighted the potentialities of the receded
barrier through its support of multiple activities such as a fishing platform, picnic
area, and panoramic views, also maintaining the scope of the barrier to be "in
line with the landscape". Other measurable co-benefits include an enhanced local
economy that will benefit from the reduced risk of inundation of the agricultural
lands behind the green barrier. However, with this solution, few agricultural lands
are expected to be floodable (Figure 6-11) and this caused some discontent
among stakeholders.

261
Figure 6-11. Aerial photo of the area with the location of the existing flood barrier and the new flood barrier
(top); Visualization of the area with the potential multiple actions that can be supported by the flood barrier
(by AgenceTer, bottom).

6.5.3 Indicators for the NBS performance assessment

The indicator matrix tailored to this demonstrator site encompasses a total of 47


indicators. Quantitative, risk-related indicators include Peak Flow volume,
Flooded Area — calculated through hydraulic modelling — and Exposed residential
and productive areas, obtained by GIS mapping. Ecosystem indicators are aimed
to assess both the effects on water quality, such as the Change in physical and

262
chemical water parameters, and water quantity, such as the Total predicted soil
loss (RUSLE), or enhanced Water storage capacity. Indicators for assessing the
improved value of the forested floodplain include Typical vegetation species
cover, and Diversity in plant and animal functional groups. Societal-related
indicators include the Number of visitors in the new recreational areas and New
pedestrian/cycling paths, whilst the Number of jobs created in the nature-based
sector is one of the economy-related indicators. The variables and key
performance indicators selected to be monitored in the Gudbrandsdalen
demonstrator site are listed in Table 6-3.

Table 6-3. PHUSICOS project key performance indicators (KPIs) to be evaluated for Gudsbrandsdalen
demonstration site.

Ambit Criterion Sub - Criterion Indicator Metric

Peak Flow m3/s


Flooding Risk
Hazard
Resilience
Flooded Area ha

Urban /Residential
ha
Areas
Potential Areas
Exposed to Risks Productive Areas
(Agriculture, Grazing, ha
Industries)

Inhabitants no./ha
RISK REDUCTION

Potential Other People


Population (Workers, Tourists, no./ha
Exposed to Risks Homeless)

Exposure Elderly, children,


no./ha
disabled

Housing no.
Potential
Buildings
Exposed to Risks Agricultural and
no.
Industrial Buildings

Roads km

Potential
Infrastructures Lifelines (Water main,
Exposed to Risks Sewerage, Pipeline,
m/km2
etc.)

263
Potential
Population
Population no.
Vulnerable to
Risks

Economic Value of the


Potential Productive Activities
Economic Effects Vulnerable to Risk (i.e. €/km2
Vulnerability due to Risks Economic Value of the
Fields, Workers No.)

Buildings No./km2
Potential
Infrastructures
Vulnerable to Transportation
Risks Infrastructures and m/km2
Lifelines

Initial costs million €


TECHNICAL & FEASIBILITY ASPECTS

Maintenance costs million €

Cost-Benefit
Analysis of the Replacement costs €
Intervention
Technical
Feasibility Avoided costs million €
(Affordability)

Payback Period years

Application of
Suitable Material used
0/1
Materials and coherence
Technologies

Physical parameters °

Effects on Water Chemical Pollution


Water -
Quality Parameters
ENVIRONMENT & ECOSYSTEMS

Water Storage
m3
Capacity Enhancement

Soil Physical Total Predicted Soil


Soil T·ha-1yr-1
Resilience Loss (RUSLE)

Typical Local
Species Typical Vegetation
Vegetation -
Promotion and Species Cover
Development

Landscape Abundance of
Green km/ha/Shannon
(Green Ecotones/Shannon
Infrastructure index
Infrastructure) Diversity

264
Diversity of Functional
Groups (Plant Shannon index
Functional Diversity)
Functional
Diversity
Diversity of Functional
Groups (Animal Shannon index
Biodiversity
Functional Diversity)

Site Community
Importance (SCI) And
Protected Areas ha
Special Protection
Areas (SPA)

Number of Visitors in
New Recreational no.
Areas

Different Activities
Leisure and Allowed in New no.
Connections Recreational Areas
Increasing
Quality of life
New Pedestrian,
Cycling and Horse m
Paths

Rate of Increase in
Social Justice %
Properties Incomes

Citizen Involved no.


SOCIETY

Stakeholders Involved no.

Participatory
Community Processes and
Involvement Partnership Public-Private
and Governance no.
Partnership Activated

Policies Set Up to
no.
Promote NBS

Social Active
Identity no.
Associations

Landscape and
Heritage
Heritage Natural and Cultural
no. of sites
Accessibility Sites, Made Available

265
Viewshed km2

Landscape
Perception
Scenic Sites and
no.
Landmark Created

Jobs Created in The


no.
Nature-Based Sector

Jobs Created in The


Nature-Based Solution
no.
Construction and
Promotion of Maintenance
Socio-
Revitalization of
Economical
Marginal Areas
LOCAL ECONOMY

Development of
Gross Profit from
Marginal Areas €/area/y
Nature-Based Tourism

Touristic Activeness
no. visitors/y
Enhancing

New Areas Made


New Areas for Available for
Local Economy Traditional Traditional Activities ha
Reinforcement Resources (Agriculture,
Livestock, Fishing, …)

6.6 Case study #3: Landslides and debris flows, Portofino Natural
Park, Italy

The Portofino Promontory (Liguria, Italy) belongs to the Natural Regional Park of
Portofino, located between Genoa and the border with Tuscany. The promontory
encompasses an area of 18 km2, with a coastal development of 13 km. The terrain
topography is rather mountainous, with high elevations over a short distance
from the coastline (e.g., Mt. Portofino with an elevation of 610 m above sea
level). Due to its unique geomorphological features, the Portofino Promontory is
historically affected by geological instabilities produced by meteorological events,
with potential impacts to the elements at risk. The most frequent hazards are (1)
shallow landslides and flash floods; (2) sea storm surges; and (3) rock falls and
mud–debris flows.

Considering the high naturalistic value of the area, NBS are the most suitable risk
mitigation measures to be adopted, to conserve landscape, natural and cultural
heritage, and touristic value of the promontory. The primary NBS ambition in San

266
Fruttuoso is to address the following challenges: (1) stabilisation of rock masses;
(2) reduction of geo-hydrologic risks in order to intercept and reduce the floating
and solid transport along the rivers and to reduce erosion; (3) wood amelioration,
by removing allochthones and degraded species of old vegetation; and (4)
construction of dry stone walls and restoration of abandoned terraces, with the
aim to valorise the terraced landscape and promote agricultural activities.

The RECONECT project foresaw the selection, installation, and operation of hydro-
meteorological instruments that will include three weather stations, two
hydrological measuring stations, and two cameras. The necessary equipment will
be purchased and installed once the selection of indicators for the evaluation of
NBS is complete. Monitoring activities further include remote sensing activities
such as LIDAR surveys, orthophotography, and infrared aerial photography.

The RECONECT project team has identified the key variables and indicators that
need to be monitored and assessed in all NBS demonstration sites. The variables
and key performance indicators selected from the original performance indicator
table to be monitored in the Portofino Natural Regional Par, are listed in Table
6-4. These assessments will be cross-referenced and compared with other
RECONECT sites that have similar morphological features (Turconi et al., 2020).

Several benefits and co-benefits are expected to be obtained from the Portofino
NBS demonstration case:

1. Decrease of geo-hydrological vulnerability for the main infrastructures and


the cultural heritage;

2. Re-building/maintenance of dry stone walls, which will contribute to the


restoration of old terraces and will re-incentivize agricultural activities with
benefits for the farmers, as well as for geo-hydrological risk mitigation;

3. Decrease of the impacts by landslides and slope instability at the coastal


sediment amount level;

4. Decrease of the risk of injuries among the park visitors due to slope
instability of interesting hiking paths during heavy rainfalls;

5. Support for the interaction between private landowners;

6. Integration of the proposed NBS with regional policies for land


management/planning and with the Basin Master Plan;

7. Improvement of the visibility and governance model of the Portofino


Natural Regional Park, also in the perspective of becoming a National Park;
and

8. Improvement of the collaboration between the park authority and


stakeholders.

267
Table 6-4. RECONECT project key performance indicators (KPIs) to be monitored in the Portofino Natural
Regional Park area (following Turconi et al., 2020).

Performance key Variable Base- Specific Monitoring Details


indicator line1
Monitoring Approach Data2 Phase3

Precipitation (mm),
Rainfall intensity Weather stations (a) 2
(mm/h)
Possible source of
debris/hyper- Assessment of terraced
concentrated flow area extent (e.g.,
Maintenance level of LIDAR), Aerial photo
(b) 2
man-made terraces interpretation and Field
survey to evaluate
terrace conditions
WATER

Floating transport in Dead trees within


20 m buffer along
hydrographical the hydro-graphical
Field survey 2
network network

Landslide reduction—
debris and hyper- Aerial photo
Land use 2
concentrated flow interpretation
triggering

Aerial photo
Changes in riparian Riparian habitat
interpretation and Field 2
habitat area (km²)
survey

Aerial photo
Changes in terrestrial Terrestrial habitat
interpretation and Field 2
habitat area (km²)
survey

Vegetation along Aerial photo


watercourses interpretation and Field 2
(survey) survey

Aerial photo
Trends and status of
interpretation and Field 2
range
Changes in vegetation survey
along watercourses
Aerial photo
Trends and status of
NATURE

interpretation and Field 2


the area
survey

Structure and Aerial photo


function including interpretation and Field 2
typical species survey

Aerial photo
Change in land cover Land cover
interpretation
1

Type of protected
Field survey 2
species
Number and type of
protected species

Number of protected
Field survey 2
species

268
Footpath network Length of improved
Field survey 2
path
recovery through
erosion reduction and
improvement of path Water drainage
smoothness Field survey 2
improvement

Increasing
Number of
recreational
recreation activity in Field survey 2
opportunities of NBS the area
area

Number of tourists Number of tourists Automatic counter 2


PEOPLE

Economic losses and


properties loss
during hydro- Survey 2
Maintenance and meteorological
management cost of events
NBS
Cultural heritage
Survey 2
loss

Maintenance and
management cost of
Survey 2
grey infrastructure
Reduced need for (if implemented)
management and
maintenance
Maintenance and
management cost of Survey 2
NBS

1)
indicates an existing baseline
2)
indicates text data; indicates vector data; indicates spreadsheet data (e.g., Excel)
3)
Number of checks in the monitoring phase: (1) represents pre- and (2) post-NBS implementation monitoring
a)
Providing data with high temporal (hourly) resolution
b)
The extent of terraced areas is only partially known as baseline

269
6.7 Case study #4: Floods in dense urban environments, Dodder
Catchment, Dublin, Ireland

This case study illustrates the case of reducing flood risk in dense urban
environments using NBS, using the example of the OPERANDUM’s OAL in the
Dodder Catchment in Dublin, Ireland. The River Dodder is one of the principal
rivers in Dublin, it flows from the Dublin Mountains through a number of high-
value dense residential areas of Dublin before discharging into the River Liffey
estuary at Ringsend (recently named “Silicon Docs” because located where all the
headquarters of the Tech Companies are). The River Dodder has a history of
flooding and is known as a river which responds quickly to a rainstorm event (Pilla
et al., 2019), mostly because of the steep gradient of the river in its upper
section. In the last century, it has overflowed its banks on numerous occasions
causing damage to adjacent properties: in 1986, when Hurricane Charlie hit
Dublin, over 300 properties surrounding the Dodder catchment were flooded (De
Bruijn and Brandsma, 2000); in February 2002, a strong high tide occurred and
over 600 properties were flooded (Javelle et al., 2002); in October 2011, a similar
number of properties were flooded throughout the catchment.

Over the past few decades, Dublin has experienced increasing pressure on land
due to population growth, urbanisation and industrialisation. The change in land
use and land cover (LULC) patterns in Dublin over the past two decades was
assessed performing both supervised as well as unsupervised classification on
LANDSAT satellite imagery data, and the effect of LULC change in streamflow
simulation was quantified by using a rainfall-runoff model (Basu et al., 2020).
Furthermore, a set of indices such as vegetation index, building index, water
index and drought index were estimated, and their changes were monitored over
time. Soil Water Assessment Tool (SWAT)-based rainfall-runoff models were used
to simulate the changes in runoff due to the LULC changes in watershed over two
decades. The results indicated an increased rainfall-runoff in Dublin due to the
high level of urbanisation, with negative impacts on flood risk in the OAL area.
This pressure is going to increase in Dublin as result of climate change in the near
future (Gharbia et al., 2016).

The high premium for land in Dublin due to the pressure on house and commercial
rental markets is resulting in less available space for the deployment of NBS to
mitigate flood risk. After a reiterative co-design approach with high level
stakeholders aimed at highlighting local challenges and drivers, and at identifying
suitable locations and typologies of interventions, the green roof was selected as
the potential NBS. The green roof has high potentials in terms of water retention,
and it could be deployed in several locations in a dense urban environment where
land has a high premium. Subsequently, rainfall-runoff-based hydrological
modelling was performed to assess the potential flood hazard areas and to
identify an effective location for implementation of NBS. For this purpose, the
hydrological model was simulated with and without the presence of NBS at
different potential locations and the site exhibiting highest flood control was
selected to be the optimal location. The selected location is in correspondence of
the CHQ building and adjacent to River Liffey, which is the main river in Dublin
(Sarkar et al., 2020). This intervention in the OAL will also be assessed through
quantitative and qualitative comparative analysis to quantify the biophysical and
economic values of different NBS alternatives and ecosystem services in Dublin

270
using two spatially explicit integrated models, Integrated Valuation of Ecosystem
Services and Trade-off (InVEST) and Soil Water Assessment Tool (SWAT), to
provide valuable data for future policies and replication of the NBS across the city
(Sannigrahi et al., 2020).

The green roof will be deployed on a roof area of around 70 m² using modular
units. The modular units will be built using exclusively recyclable materials. In
order to demonstrate the effectiveness of the green roof NBS, some of the
modular units will be left empty without any soil and vegetation: this will allow
to assess the performances of the vegetated units in terms of water retention
during the pilot time. The assessment will be carried out by instrumenting the
green roof with a dense network of sensors. Specifically, the following sensors
will be deployed: (1) rain gauges to measure rainfall; (2) sensors to measure
wind speed/direction, humidity, temperature, and solar radiation; (3) soil
moisture sensors for the piloted modular units; (4) rain gauges to measure the
water exiting the modular units; and (5) cameras to visually monitor the green
roof and create time-lapse videos for engagement activities. A dashboard with
the sensors data and the time-lapse will be displayed on a screen in the CHQ
shopping centre to increase the public awareness on the green roof NBS and its
potential to reduce flood risk. The concept behind this solution is to bring nature
online as the next frontier in ecosystem management with the aim to change the
relationship with the natural world in an age of rapid urbanisation and digitisation
(Galle et al., 2019).

The framework developed by the OPERANDUM consortium for vulnerability and


risk assessment of social-ecological systems (SES) subjected to natural hazards
will then be utilised to more comprehensively assess the green roof intervention
(Shah et al. 2020), with the aim to provide the City Council with valuable
information for future policies and thus foster the replication of the NBS piloted
in the OAL in Dublin. The detailed smart green roof approach will then be
replicated on other public buildings owned by Dublin City Council to further
mitigate flood risk in the dense urban environment of Dublin city.

Finally, the OAL activities related to the assessment and wider deployment of the
green roof NBS include the spatial reconfiguration and optimisation of the dense
network of rainfall sensors (over 50) in the Dublin area. This is done with the
support of Dublin City Council who provided access to the sensors. The statistical
models used for this task replicate and expand the work detailed in Basu et al.
(2019), which allows the identification of redundant rain gauges and influential
ungauged locations in the Greater Dublin area based on hourly and daily rainfall
data by considering covariance factor, kriging, Shannon entropy and annealing
approaches. The data from the optimised network of rain gauges will be then
used, in conjunction with the measurements from the river level sensors, to
generate Artificial Intelligence forecasting models for river levels, which will allow
to alert the Council of potential flood events according to different weathers,
replicating an approach used previously in another Irish catchment (Assem et al.,
2017).

271
6.8 Concluding Remarks

Effective disaster risk reduction strategies require a combination of several


techniques, and implementation of structural and non-structural measures.
Choosing the optimal strategy is a key objective for local authorities and
infrastructure managers.

NBS can be considered as structural measures with sometimes limited capacity


(for mitigating the impacts of extreme events, for example) but also with
additional co-benefits in comparison with classical grey measures. Needless to
say, no solution can be universal and work in all situations. NBS may exhibit some
drawbacks: during extreme floods, riparian forests supply woody debris which
worsen the risk level. Most of time, a hybrid combination of green and grey
measures will provide the optimal solution when DRR is the main goal (e.g.,
riparian buffers and a rack to trap large debris just upstream sensitive bridges).
NBS assessment requires consideration of several criteria and combined
methods. Assessment frameworks based on classical deterministic approaches
cannot be used alone anymore. Other frameworks such as decision-aiding
methods and systemic analysis offer new opportunities and methodologies. A
paradigm shift in DRR engineering is probably emerging through the recent NBS
projects (see Tacnet et al., 2019).

To assess the effectiveness of any measure, the analyst must identify its function,
the required capacity of the measure being assessed and a measurable indicator
for evaluating this capacity. Classical indicators used for risk assessment can be
employed for this purpose. The case studies provided here are only partial
examples and should be considered more as non-exclusive methodological
pathways to characterize NBS effectiveness. The fact that NBSs are effective for
mitigation of the impacts of extreme events has still to be demonstrated. To
mitigate the risk of extreme natural hazard events, classical civil engineered
techniques and hybrid solutions may be the optimal measures in the foreseeable
future.

Finally, a DRR strategy based on NBS faces the same large challenges linked to
any DRR strategy, including multi-risk situations, global change effects and
uncertainties.

6.9 References

Alves, A., Gersonius, B., Kapelan, Z., Vojinovic, Z., and Sánchez, A., ‘Assessing the Co-Benefits of green-blue-
grey infrastructure for sustainable urban flood risk management’, Journal of Environmental
Management, Vol. 239, 2019, pp. 244–254.

Alves, A., Vojinovic, Z., Kapelan, Z., Sánchez, A., and Gersonius, B., ‘Exploring trade-offs among the multiple
benefits of green-blue-grey infrastructure for urban flood mitigation’, Science of the Total
Environment, Vol. 703, 2020.

Arfaoui, N. and Gnonlonfin, A., Testing Meta-Regression Analysis in the context of NBS restoration measures:
The case of Brague River, Working Paper ESDES n°2020-02, 2020a. Available from:
https://www.esdes.fr/wp-
content/uploads/sites/12/2020/11/wp_esdes_2020_02_arfaoui_gnonlonfin.pdf

272
Arfaoui, N., and Gnonlonfin, A., ‘Supporting NBS restoration measures: A test of VBN theory in the Brague
catchment’, Economics Bulletin, Vol. 40, No 2, 2020b, pp. 1272–1280.

Assem, H., Ghariba, S., Makrai, G., Johnston, P., Gill, L. and Pilla, F., ‘Urban water flow and water level
prediction based on deep learning’, In: Joint European Conference on Machine Learning and Knowledge
Discovery in Databases, 2017, pp. 317-329.

Basu, B., Sarkar, A., and Pilla, F. ‘Identification of optimal number of rain gauges and their locations based on
different statistical approaches: A case study in Dublin based on hourly and daily rainfall data’,
Geophysical Research Abstracts, 2019.

Basu, B., Sarkar Basu, A., Sannigrahi, S., and Pilla, F., ‘Investigating land use and land cover changes in
Dublin, Ireland using Satellite Imagery: A comparative analysis’, In: EGU General Assembly 2020,
Online, 4–8 May, 2020.

Bokhove, O., Kelmanson, M.A., Kent, T., Piton, G., and Tacnet, J.-M., ‘Communicating (nature-based) flood-
mitigation schemes using flood-excess volume’, River Research and Applications, Vol. 35, 2019, pp.
1402–1414.

Bokhove, O., Kelmanson, M.A., Kent, T., Piton, G., and Tacnet, J.-M., ’A Cost-Effectiveness Protocol for Flood-
Mitigation Plans Based on Leeds’ Boxing Day 2015 Floods’, Water, Vol. 12, 2020, pp. 1–30.

De Bruijn, E., and Brandsma, T., ‘Rainfall prediction for a flooding event in Ireland caused by the remnants of
hurricane Charley’, Journal of Hydrology, Vol. 239, 2000, pp. 148–161.

Evette, A., Labonne, S., Rey, F., Liebault, F., Jancke, O., and Girel, J., ‘History of Bioengineering Techniques
for Erosion Control in Rivers in Western Europe’, Environmental Management, Vol. 43, 2009, pp. 972-
984.

Galle, N.J., Nitoslawski, S.A., and Pilla, F., ‘The Internet of Nature: How taking nature online can shape urban
ecosystems’, The Anthropocene Review, Vol. 6, No 3, 2019, pp. 279-287.

Gharbia, S.S., Gill, L., Johnston, P., and Pilla, F., ‘Multi-GCM ensembles performance for climate projection on
a GIS platform’, Modeling Earth Systems and Environment, Vol. 2, No 2, 2016, p. 102.

Hellmers, S., Ackermann, D., Einfalt, T., and Fröhle, P., ‘Konzeptstudie zur Steuerung von
wasserwirtschaftlichen Anlagen auf der Grundlage von Ensemble Kurzzeitvorhersagedaten’, In: Tag
der Hydrologie, Trier, Germany, 2017.

Javelle, P., Ouarda, T.B., Lang, M., Bobée, B., Galéa, G., and Grésillon, J.-M., ‘Development of regional flood-
duration–frequency curves based on the index-flood method’, Journal of Hydrology, Vol. 258, 2002,
pp. 249–259.

Pengal, P., Pagano, A., Piton, G., Kozinc, Z., Cokan, B., Šinkovec, Z., and Giordano, R., ‘Chapter 16: Glinščica
for all: exploring the potential of NBS in Slovenia: barriers and opportunities’, In: WaterSecurity in a
New World, Springer, 2020.

Pilla, F., Gharbia, S.S., and Lyons, R., ‘How do households perceive flood-risk? The impact of flooding on the
cost of accommodation in Dublin, Ireland’, Science of The Total Environment, Vol. 650, 2019, pp.144-
154.

Rinaldi, M., Surian, N., Comiti, F., and Bussettini, M., ‘A method for the assessment and analysis of the
hydromorphological condition of Italian streams: The Morphological Quality Index (MQI)’,
Geomorphology, Vol. 180-181, 2013, pp. 96–108.

Sannigrahi, S., Basu, B., Sarkar Basu, A., and Pilla, F., ’Ecosystem service-based approach for evaluating the
effectiveness of nature-based solution in mitigating climate change and land degradation issues in a
city region’, EGU General Assembly 2020, Online, 4–8 May, 2020.

Sarkar Basu, A., Basu, B., Sannigrahi, S., and Pilla, F., ‘Deployment of Green roof top as a Nature Based
Solution in Dublin, Ireland’, EGU General Assembly 2020, Online, 4–8 May, 2020.

Shah, M.A.R., Renaud, F.G., Wild, A., Anderson, C.C., Loupis, M., Panga, D., and Sabatino, S.D., ‘A conceptual
framework for vulnerability and risk assessment in the context of nature-based solutions to hydro-
meteorological risks’, EGU General Assembly 2020, Online, 4–8 May, 2020.

Tacnet, J.-M., Piton, G., Philippe, F., Gourhand, A., and Vassas, C., ‘Décider dans le contexte de la GEMAPI :
exemple de méthodologie d’une approche intégrée d’aide à la décision et application aux projets
d’aménagements’, Science Eaux and Territoires, Vol. 26, 2018, pp. 48–53.

Tacnet, J.-M., Piton, G., Favier, P., Pengal, P., DELIVERABLE 5.4 Integrative modelling framework and testing
in the DEMOs. Part 4: From indicators definition to NBS choice and effectiveness assessment, EU
Horizon 2020 NAIAD Project, Grant Agreement N°730497, 2019.

273
Turconi, L., Faccini, F., Marchese, A., Paliaga, G., Casazza, M., Vojinovic, Z., and Luino, F., ‘Implementation
of nature-based solutions for hydro-meteorological risk reduction in small Mediterranean catchments:
The case of Portofino natural regional park, Italy’, Sustainability, Vol. 12, No 3, 2020, p. 1240.

Watkin, J. L., Ruangpan, L., Vojinovic, Z., Weesakul, S., and Torres, S.A. (2019). ‘A Framework for Assessing
Benefits of Implemented Nature-Based Solutions’, Sustainability, Vol. 11, No 23, 2019, p. 6788.

274
MAES
Mapping and Assessment of Ecosystems and their Services
Urban pilot and EU ecosystem assessment
EU GB

Action 5 of the Strategy, better known as Mapping and Assessment of Ecosystems and their Ser-
vices (MAES), states ‘Member States, with the assistance of the Commission, to map and assess
the state of ecosystems and their services in their national territory, assess the economic value
of such services, and promote the integration of these values into accounting and reporting sys-
tems at EU and national level by 2020’. MAES provided guidance to EU countries on ecosystem
assessment through a series of thematic pilots including urban ecosystems. It also delivered a
EU ecosystem assessment which provides an analysis of trends in pressures, condition and ser-
vices of marine, freshwater and land ecosystems of the EU+GB using 2010 as baseline year. Ur-
ban ecosystems cover about 5% of the EU land area but their immediate impact stretches well
beyond their boundaries. Therefore, the system of functional urban areas, which cover 22.5%
of the EU land area, was used in the assessment to analyse trends in pressure and condition.

Average greenest (2010)


within Urban GI
Commuting zone - densely built
Reykjavik <0.3
0.3-0.4
0 125 Km 0.4-0.45
0.45-0.5
60°N

> 0.5

Helsinki Population size


(in core cities)
Oslo Tallinn
Stockholm < 200000
20000-500000
Riga
500000 - 1500000

København Vilnius > 1500000


Dublin
EU average: 0.45
50°N

Berlin Warszawa
Amsterdam
London

Brussel

Praha
Luxemb.
Paris
Kishinev
WienBratislava
Budapest

Bern

LjubljanaZagreb
Bucuresti
Beograd

Sarajevo
Sofiya
Podgorica
Skopje
40°N

Roma Tirana
Madrid

Lisboa

Athinai

Nicosia

El-Jazair
Tounis
0 250 Km
Valletta
0° 20°E

Drawing on knowlegde from projects


funded by the European Union
Figure: Average greenest (2010) within Urban Green Infrastructure - Maes, J. et al. 2020
SCOPE EU wide ecosystem assessment

Approach to Impact Assessment Main Challenges addressed

Urban ecosystems, cities and their surroundings


were assessed using the functional urban areas 1. Climate Resilience
implementing the MAES framework to assess
ecosystems pressures and condition. Indicators
2. Water Management
are spatially explicit and were implemented in a
3. Natural and Climate Hazards
consistent and comparable way.
4. Green Space Management

5. Biodiversity
Involved Stakeholders and roles
6. Air Quality

The MAES initiative is a collaboration between 7. Place Regeneration


three key stakeholder groups:
8. Knowledge and Social Capacity Building
• the member states, represented by natio-
nal environmental authorities; 9. Participatory Planning and Governance
• EU services including DG Environment, DG
Research and Innovation, the Joint Re- 10. Social Justice and Social Cohesion
search Centre, and the European Environ-
ment Agency; 11. Health and Wellbeing
• scientists and experts who participate to
the working group or who contributed to 12. New Economic Opportunities & Green Jobs
pilots and case studies.

Nationale environmental authorities

EU level stakeholders

Scientists / Academia

Lessons learned

Growing cities can minimise their impacts


through nature-based solutions. These solutions
span from the conservation of natural ecosys-
tems within commuting zones to the restora-
tion, creation and management of multifunctio- Learn more
nal green urban areas in order to improve local on the project website
climate, reduce urban overheating, mitigate
flooding, air pollution and biodiversity loss. The
design and management of new urban nature in
the core area of cities can also provide opportu- Learn more
nities for recreation and social interaction, and in the final Report
significantly improve urban quality of life.

Maes, J. et al. Mapping and Assessment of Ecosystems and their Services: An EU ecosystem assessment,
EUR 30161 EN, Publications Office of the European Union, Luxembourg, 2020, ISBN 978-92-76-17833-0 (on-
line),978-92-76-22954-4 (supplement), doi:10.2760/757183 (online),10.2760/519233 (supplement), JRC120383.

MAES is a JRC Institutional project


EnRoute
Enhancing Resilience of Urban Ecosystems
through Green Infrastructure
Antwerp (BE) Dublin (IE) Glasgow (GB) Helsinki (FI) Karlovo (BG) Leipzig (DE)
Lisbon (PT) Manchester (GB) Oslo (NO) Poznan (PL) Padova (IT) Rome (IT)
Tallinn (EE) The Hague (NL) Trento (IT) Utrecht (NL) Valletta (MT)

EnRoute is a project of the European Commission in the framework of the EU Biodiversity


Strategy and the Green Infrastructure Strategy. EnRoute provides scientific knowledge of
how urban ecosystems can support urban planning at different stages of policy and for vari-
ous spatial scales and how to help policy-making for sustainable cities.

Drawing on knowlegde from projects


funded by the European Union
Image: Seurasaari Island, Helsinki - Photo © Leena Kopperoinen
SCOPE How urban ecosystems can support urban planning

Approach to Impact Assessment Main Challenges addressed

EnRoute (2017-2019) was an Administrative Agree-


1. Climate Resilience
ment between DG Environment and the JRC, with a
view to implement a European Parliament Pilot Pro-
ject on Urban Green Infrastructure. This Pilot Pro- 2. Water Management
ject aimed at building further on the many positive
experiences of the MAES urban pilot, for which the 3. Natural and Climate Hazards
JRC has been a central partner. It will further help
promoting the application of GI at local level and 4. Green Space Management
will deliver guidance on the creation, management
and governance of GI, e.g. through the develop- 5. Biodiversity
ment of relevant indicators to map and assess GI
in urban contexts, and testing them in additional ci- 6. Air Quality
ties. The approach was built on three lines of focus:
(1) Networking and improving flows of knowledge 7. Place Regeneration
and information, (2) Citylabs: Testing the urban GI
MAES framework in cities across Europe. (3) Better 8. Knowledge and Social Capacity Building
understanding of positive interactions and mutual
reinforcement between urban GI policies and objec- 9. Participatory Planning and Governance
tives at different governance levels. EnRoute also
contributed to the impact evaluation framework for 10. Social Justice and Social Cohesion
nature-based solutions (EKLIPSE project) and pro-
vides tools and protocols for measuring the impact 11. Health and Wellbeing
of urban nature-based solutions.
12. New Economic Opportunities & Green Jobs
Involved Stakeholders and roles

18 city-labs were involved. A research institute be scaled-up further if we are to create more re-
and the municipality were involved. City-labs im- silient, sustainable and ‘livable’ cities for future
plemented the MAES approach to assess urban generations. This project provided knowledge on
ecosystems and urban GI, focusing on challenges how UGI can support urban policy-objectives at
discussed with the local authorities. different stages of the planning process and at a
variety of spatial scales.
Municipal Administrations
The proposed framework is useful for
• Making a case at local level
Scientists / Academia • Compare the performance of cities
• Raising awareness about the multiple functio-
nality of ecosystems
Lessons learned • Enhancing cross-sector cooperation or co-
operation across different political levels
Urban Green Infrastructure (UGI) refers to the
strategically managed network of urban green EnRoute:
spaces and natural and semi-natural ecosystems • Provided inspiration at national and local level
situated within the boundary of an urban ecosys- • Provides a framework that can be adapted to
tem. These high-quality, biodiversity-rich areas fit local needs
can help make cities more sustainable and con- • Helped build communities of practice across
tribute to solve many challenges, such as air pol- sectors
lution, noise, climate change impacts, heat wa- • Provided diverse set of examples/city-labs
ves, floods and public health concerns. As cities gives inspiration
grow and develop, it is vital to improve the avai-
lability, quality and accessibility of UGI. Urban
planners and decision-makers across Europe are
increasingly seeking to integrate UGI, ecosystem Learn more
services and nature-based solutions into their https://oppla.eu/groups/enroute
urban planning processes, but these efforts must

EnRoute (2017-2019) was an Administrative Agreement between DG


Environment and the JRC, with a view to implement a European Parliament
Pilot Project on Urban Green Infrastructure.
7 DATA REQUIREMENTS
Coordinating Lead author
Leo, L.S.

Lead authors
Kalas, M., Leo, L.S.

Contributing authors
Baldacchini, C., Budau, O.E., Castellar, J., Comas, J., Connop, S., Corbane, C., Decker, S.,
Draghia, M., Dubovik, M., Dushkova, D., Haase, D., Ivits, E., Körmöndi, B., Kumar, P., Laikari,
A., Leopa, S., Littkopf, A., Ommer, J., Rinta-Hiiro, V., Spano, G., Spinnato, P., Vranić, S.,
Teixeira da Silva, R., Zavarrone E.

Summary

What is this chapter about?

Chapter 7 offers an overview of the main types of data, data sources, and data
generation techniques for NBS monitoring and impact assessment. After
familiarising you with common data terminology and definitions (Section 7.1), we
review the types of data associated with NBS monitoring and assessment
(Sections 7.2–7.7), their use for indicator assessment (Section 7.8) and baseline
construction (Section 7.9), and the principal aspects determining the quality of
analysis (Section 7.10). Concepts are illustrated through examples and
complemented with potential data sources. Finally, we reflect on data sharing,

279
data exchange, data management and dissemination of data gathered (Section
7.11).

How can I use this chapter in my work with NBS?

This chapter aids to understand the data requirements for evaluating NBS
performance and impact. This chapter:

1) Provides knowledge regarding available data sources;

2) Assists in developing a robust plan for the collection, management and use of
data;

3) Offers examples of how data have been collected and integrated by various
EU Horizon 2020 projects; and,

4) Raises awareness of the challenges commonly encountered such as data gaps,


data availability, data reliability and related potential error sources.

When should I use this knowledge in my work with NBS?

The knowledge provided in this chapter can be used in the planning phase of NBS
projects in order to assess whether the required datasets can be obtained from
external data sources or should be generated within the project. In the latter
case, Chapter 7 provides guidance towards data generation/integration (e.g.,
modelling, measurement campaigns). This chapter also supports the
development of standardised data management protocols for effective data
sharing and data dissemination.

How does this chapter link with the other parts of the handbook?

Chapter 7 supports the development and execution of a robust monitoring and


evaluation plan (Chapters 2 and 3), by detailing considerations related to data
types, data integration, and the adequacy of data for indicator assessment and
baseline construction. This chapter describes the data requirements for
computing NBS indicators (Chapters 4-6 and Appendix of Methods).

Evaluating NBS benefits, co-benefits, and trade-offs can be a data intensive


process. Understanding the data requirements is a critical element in relation to
ensuring both the efficacy and cost-effectiveness of this evaluation process. In
order to establish the monitoring plans and schemes described in previous
chapters, and to deliver this over the range of relevant scales, it is therefore
critical to generate data that are both applicable for the nature-based solution
impact assessment, and that are comparable to the preceding monitoring
campaigns. This chapter addresses the data requirements involved in evaluating
the impacts that nature-based solutions manifest and explains the data building
blocks involved in NBS monitoring and assessment procedures.

280
Figure 7-A. How can we generate data for NBS monitoring and evaluation?

7.1 Data terminology, definitions and key concepts

Data requirements for NBS monitoring and assessment span multiple and diverse
data types and sources, and thus involve techniques, methods and concepts
drawn from various disciplines of both natural and social sciences. This section
provides the reader with a basic knowledge of the terminology and concepts
commonly encountered when dealing with data requirements for the NBS
evaluation process. It also contains explanations of the main data types and data
aspects relevant for NBS assessment and thus aids the reader in navigating the
rest of this chapter.

7.1.1 Spatial versus non-spatial data

Spatial data is a term used to describe data containing information about a


specific location on the Earth's surface. Spatial data are essential for any mapping
activity as they provide information on the exact location, shape, size, and
orientation of a given entity (e.g., a river). Non-spatial data, on the contrary,
contain information which is independent from any geometric and/or topological
consideration (e.g., street names). Non-spatial data are also termed attributes
as they are usually combined with spatial data to provide additional information
on the specific geographic entities identified by a spatial dataset. For example,
the geometric characteristics of a city district (spatial data) can be combined with
information on air quality (non-spatial data) and displayed together on a map
using a legend of colours, with each colour indicating a certain level of air
pollution.

281
Spatial data are stored in spatial databases that are optimized for storing and
querying data that represent objects defined in a geometric space. Depending on
the way they are manipulated and stored, spatial data can be of two types: vector
and raster. In vector form, spatial data are represented in form of points (e.g.,
the location of individual trees in a city), segments (e.g., the path of a river in
the same city) and polygons (e.g., houses and urban green parks). In the
simplest form of a raster, spatial data are represented as a matrix of cells (or
pixels) organized into rows and columns (a grid) where each cell contains a value
representing information (such as elevation, temperature, number of people).
Satellite images, such as land cover/land use maps, are typical examples of raster
spatial data. Manipulation, storage, and visualization of digital spatial and non-
spatial datasets are commonly done using GIS (Geographic Information Systems)
software like ArcGIS. Examples of spatial and non-spatial data of relevance for
NBS monitoring are given in Section 7.8.

7.1.2 Baseline data

As defined by EUROSTAT, a baseline study is “an analysis of the current situation


to identify the starting points for a programme or project. It looks at what
information must be considered and analysed to establish a baseline or starting
point, the benchmark against which future progress can be assessed or
comparisons made.” (EUROSTAT, 2014). In the context of NBS, the
establishment of a baseline involves collecting a set of data that allows the
description of the geo-morphological, socioeconomic conditions, living standards
and livelihoods of NBS project-affected communities and their potential hosts
prior to any NBS intervention. Those data will be used as a reference for
monitoring the impacts of the NBS on the involved territories, thus allowing a
comparison between the pre-
project implementation state of
play and the post-project
implementation situation. The
results of this monitoring process
are the starting point not for the
comparison between the changes
occurred due to NBS interventions
and other grey or hybrid solutions
addressing the same issue, but for
the assessments of the benefits
attributable to NBS. Baseline data collection and requirements are the topic of
Section 7.9.

7.1.3 Control data

Impact evaluation mostly addresses the cause-and-effect questions and different


methods can be used to establish what the causal effect (impact) of an NBS
intervention on an outcome of interest is. These methods should estimate the so-
called counter-factual: is a given NBS intervention effective compared to the

282
absence of the intervention or to alternative, traditional engineering or planning
solution? Control data are generally collected to assess counter-factual, and they
consist in collecting the same variables, with the same methodology, as per the
NBS intervention site, in a suitable, different site. Depending on the outcome to
be evaluated, control data collection would need the identification of a suitable
control area or control group. Further details on this aspect can be found in
Chapters 2 and 3 of this Handbook.

7.1.4 Acquisition regime

Acquisition regime refers to the temporal interval over which a certain variable
(e.g., temperature) or process is monitored. Typically, the timestamp assigned
to a data point can refer to discrete observation/model time (which represents
the sampling frequency) or the beginning or the end of the
observation/aggregation time interval. Following the INSPIRE Directive (EC,
2007), acquisition regime can be distinguished into:

• Continuous data acquisition (Data are generated on a continuous basis)

• Demand driven data (Data are generated on demand)

• Once-off data (Data are generated only once in this configuration. No


further observations in this configuration can be expected)

• Periodic data collection (Data are generated at regular intervals)

For example, relevant indicators such as residential property sale and rent value
in the areas of future NBS implementation, can be solely available as once-off
data. On the contrary, many of the datasets employed for baseline conditions
characterisation (cf. Section 7.9) are typically retrieved from national statistics
organisations or local municipalities, thus they have varying periodicity: at
national level they are usually collected with a yearly periodicity, while at
neighbourhood level, data collection is only done during national censuses, which
are conducted every 5 to 10 years.

In many cases, data for the computation of NBS environmental indicators are
acquired continuously, either as part of permanent monitoring networks
established by environmental agencies and research institutions or as ad-hoc
monitoring campaigns carried out within NBS projects. In the EU-H2020 project
UNaLab, for example, continuous data collection has been used for quantifying
physicochemical indicators, such as discharge and water quality, as well as for
other environmental constituents (e.g., temperature, precipitation, and air
quality).

In general, the choice of a certain acquisition regime over another should be


dictated by (and lower than) the expected temporal dynamics of the process or
variable under scrutiny. In practice, however, it is often a compromise between
several factors, such as technological feasibility, project duration, resources, and
funding availability. This means that adequate acquisition regimes should be

283
carefully assessed to avoid data gaps, poor data adequacy (cf. Section 7.10), and
limited data availability in the computation of NBS performance indicators as well
as the establishment of a baseline.

7.1.5 Spatial scale of analysis

Spatial domain is another critical factor affecting data representativeness and


adequacy. Data requirements in terms of spatial domain depends on a
combination of (1) the scale of nature-based solution intervention (large vs. small
scale NBS), and (2) the expected scale of the impact for each indicator being
evaluated (some datasets are representative of small-scale processes while
others provide impact at broader scales). This means that NBS evaluation
indicators need to be assessed over the proper spatial scales. Those can be
identified with the aid of other types of indicators which have been created with
the specific purpose of measuring the spatial scale of NBS impacts (for example,
the spatial extent of cooling effect in relation to reduced air temperature).

Thus, scale classification in terms of data requirements may include:

• Landscape or regional scale

• City scale

• Neighbourhood scale

• Street or pedestrian scale

• Nature-based solution footprint scale

The typical NBS scales involved are relatively small, namely data requirements
are usually at the neighbourhood scale, the street or pedestrian level, and the
NBS footprint scale. Nevertheless, datasets at larger scales become important
when assessing the upscaling and replication potential of individual NBS
interventions at city scale or at landscape/watershed scale (as in the case of NBS
for disaster risk reduction – cf. Chapter 6) and, in that respect, they allow to
establish robust baselines to guide planning and city-wide interventions.

An example is the series of NBS eco-gardens being implemented in kindergartens


across the city of Poznan in the framework of the EU-H2020 project Connecting
Nature. In terms of scales, the transition from hard impermeable surfaces, like
asphalt, to vegetated surfaces, is expected to positively impact the thermal
comfort at the scale of the kindergarten footprint. However, if a critical mass of
nature-based solutions can be rolled out across the city in future, through the
implementation of eco-gardens in social spaces and other mechanisms, it might
be worth considering also the establishment of a baseline for thermal comfort at
greater regional/administrative scale, so that changes compared to the baseline
can be quantified in future.

284
For ease of comparison between indicators within a location, for ease of
comparison of an indicator between cities, and in relation to exploiting data
sources that are already collected, using standardised spatial scales can be
beneficial. For example, Nomenclature of Territorial Units for Statistics (NUTS)
spatial scales for indicator evaluation can provide a standardised scale
(EUROSTAT, 2020). NUTS represent a geocode standard, developed and
regulated by the European Union, for referencing the subdivisions of countries for
statistical purposes. For EU member countries, a hierarchy of three NUTS levels
was established, corresponding to increasing granularity of districts. Whilst not
always corresponding to administrative divisions within a country, the NUTS
spatial scales correspond with standardised data gathering and reporting that can
be a useful data source for evaluation indicators, particularly those associated
with economic evaluation. It is, however, important to note that NUTS scales will
not be relevant for all expected spatial scales of impact.

7.1.6 Processing level

From a data processing perspective, the computation of a given NBS indicator


consists of using existing data to create new types of data through some sort of
transformation, such as an arithmetic formula or aggregation (e.g.,
spatial/temporal interpolation). Various degrees of data integration and
manipulation are possible, which means that basic indicators can be used as input
data for the computation of more sophisticated or synthetic indicators. In that
respect, a straight indicator hierarchy has been recently proposed within the EU-
H2020 Project Nature4Cities. This hierarchy classifies the indicators into three
levels of processing (Figure 7-1). A 1st level indicator is a value derived from a
dataset, which describes the state of a phenomenon or the environment. If a 1st
level indicator is introduced into an equation or model, it gets into the next level
which is the 2nd level indicator. If this one is used again in an equation or model,
then it is a 3rd level indicator. For each new level, assumptions are made and
accumulated, and simplification or loss of quality may result.

Figure 7-1. Indicator Hierarchy adopted in the EU-H2020 project Nature4Cities.

Depending on the specific indicator and the temporal and spatial scales under
consideration, some 1st, 2nd or even 3rd level indicators can be readily available
from external data sources (e.g., national statistics organisations and
environmental agencies). In most cases, however, the computation of NBS

285
performance indicators entails the acquisition of the required datasets. These
datasets can be retrieved from external databases (when available) or newly
generated by conducting ad-hoc measurement campaigns and/or numerical
modelling efforts (cf. Sections 7.2–7.6). In both cases, it is important to recognize
that data themselves undergo different level of processing before becoming
directly usable by non-technical experts. For example, satellite data such as
Sentinel products are systematically provided at various processing levels.

In general, there are three levels of processing commonly encountered with any
type of dataset:

• Raw data, namely data directly outputted by a measuring device or a


numerical model (or any other data acquisition technique), with no (or
minimal) data validation/verification, manipulation, or conversion into
standard units and/or formats. These data are rarely usable by non-
expert users.

• Quality controlled data, namely data which have been screened for
outliers and other possible error conditions. Data points identified as
problematic and erroneous are removed or flagged.

• Final data products, namely data which have been quality checked and
have undergone various post-processing procedures to be converted into
more useful parameters and data formats.

7.1.7 Data Generation and Collection Methods

Data collection should be based on solid planning, technical expertise, and a wide
knowledge of the state of the environment and its functioning in relation to
humans in order to ensure that the relevant and accurate data are garnered
properly for the purpose of NBS monitoring and assessment. In general, data
collection methods (also referred to as acquisition mode) used for NBS monitoring
and assessment include a few standard ways of collecting data: (a) Observations,
(b) Surveys and Census, (c) Laboratory Experiments.

Observations can be regarded as one of the main methods for monitoring the
performance of NBS interventions and their impact on the socio-ecological
system. This includes manual or automated collection of quantitative information
(namely direct measurements, e.g., measurement of temperature) or can be
defined as a detailed examination by watching, noticing or hearing (Kawulich,
2012) in case of qualitative information. Differently from survey, the observer
does not influence the study in any way or attempt to intervene in it. As such,
one of its advantage is the objectivity. In the rest of this chapter, observational
data are differentiated into population observations and environmental
observations due to their different techniques in data acquisition. For example,
satellite and ground sensor observations are primarily used for environmental
monitoring and further discussed in Sections 7.2.1 and 7.2.2, respectively.

286
On the other hand, people’s behaviour and attitude towards NBS interventions
can be also observed by other humans without direct interaction as explained in
Section 7.3.2. Population observations function as an umbrella for different
methods of collecting data on people’s behaviour, attitudes, and, especially, their
interaction with each other but also with nature. These methods have been
increasingly used for monitoring social benefits of NBS. In this context,
observations can be either quantitative (e.g., number of people visiting an NBS)
or qualitative (e.g., how people interact with nature or an NBS).

Surveys and Census represent another important method of collecting


environmental, socio-demographic and economic data and statistics for NBS
assessment. An important source of survey data for NBS are administrative
records, namely administrative data stored by the governments and other
organizations such as annual reports on the state of environment, etc.

Differently from observations, surveys represent a research strategy to collect


information in interaction with people (Ponto, 2015). Survey data are collected
by having participants (sample group or population) responding to quantitative
and/or qualitative questions. The responses of the sample group are statistically
analysed and can be used as representative, under specified conditions, of a
whole and for comparison. Census data differ from (quantitative) survey data
only in terms of completeness and for temporal slices. Indeed, while survey data
are based on a population sample, census data are universal by considering every
individual. In regard to NBS, survey data can be used for defining a baseline and
for further monitoring of socio-economic and health benefits and impacts. Section
7.3.1 addresses survey data in more detail.

It should, however, be noticed that the term survey is also frequently used in the
context of environmental monitoring, mainly to indicate data collection methods
which require sampling (e.g., removal of the soil) of the object of investigation.
This type of survey is for example used to monitor biodiversity at the NBS site
(cf. Section 7.2.3).

Laboratory Experiments are useful when the researchers intend to control the
results of the study always in a cause and effect pattern (Sullivan et al., 2016).
Differently from observational studies which randomly select a sample and may
find correlations between variables (Rosenbaum, 2010), laboratory studies can
control or manipulate some or all variables that might affect the phenomenon
under study and thus identify and confirm the potential mechanisms underlying
observed responses (Montgomery, 2008). In the context of NBS, laboratory
studies can help assessing either people behaviour towards NBS or the
environmental performances of different NBS. In either application, laboratory
data could be particularly valuable when used as pilot studies and/or at the
planning phase of an NBS intervention, as discussed in Section 7.6.

Data collected through the aforementioned methods are typically complemented


by data generated through modelling approaches. Numerical simulations and
modelling refer to a fundamental part of the methodologies used in NBS
monitoring and will be discussed in Section 7.5. Modelling is a process of
abstraction and generalization aimed at developing adequate models
(representations) of the real-world systems to be examined (Grützner, 1996).

287
The models developed for data simulation purposes can be classified as simplified
(non-physics-based) model and numerical models (although other categories and
classifications exist). Simplified conceptual models are a representation of
physical processes and require significantly less computer effort than the
numerical models. They are particularly appropriate to simulate datasets for large
study areas and/or stochastic modelling for probabilistic based risk assessment
(including elements of randomness, e.g., probability distributions and generalised
linear models) and multi-scenario modelling on a bigger scale with availability of
quality observational datasets. Numerical models are mathematical equations
that attempt to simulate a state variable by solving equations developed by
applying laws of physics and typically require solving them computationally.
Therefore, the numerical models are developed to represent/simulate detailed
state variables (e.g., temperature, precipitation) dynamics. Depending on their
spatial representation of the problems in hand, the models use lumped (variables
of interest are a function of time only) or spatially distributed approach and can
be dimensionally classified into one-dimensional (1D), two-dimensional (2D) and
three-dimensional (3D) models.

Citizen Science is a research focus that enables citizens and stakeholders to be


actively engaged in science data generation and monitoring programs. It refers
to “the general public engagement in scientific research activities when citizens
actively contribute to science either with their intellectual effort or surrounding
knowledge or with their tools and resources” (Serrano et al., 2014). The European
Environment Agency define three types of citizen science activities based on the
degree of citizen involvement: 1. contributory – meaning that citizens are
involved in data collection; 2. collaborative - participants are involved in more
than data collection such as in data analysis, project design, and results
dissemination; and 3. co-created – where citizens are involved in basically every
aspect of work. Citizen science opens new possibilities for data collection and
analysis, introduces different perspectives and cooperation, but also offers
various benefits for the community itself, such as public engagement, awareness
raising, and lifelong learning opportunities in science (Hecker et al., 2018). In
terms of data generation, citizen science can generate a range of different data
types. This approach is primarily used for environmental monitoring, but there
are also examples of social and economic applications. Citizen Science has also
been increasingly used in NBS context. This will be discussed in Section 7.7.

Another emerging approach is Big Data. The term indicates data which are
characterized by large variability, volume and variety, among other aspects. Big
data can be considered as an evolution of “data mining”, which refers to the
development of datasets which are very large and can be identified with statistical
significance (Sang, 2020). Data mining means searching for valuable information
in a large database. Deploying data mining methods requires a type of expertise
which is increasingly in demand, but this expertise is not domain-specific. It can
be deployed where scientific theory has no more intelligent solution to offer
(Sang, 2020). Despite the several pitfalls hidden into it, the use of big data could
be key in the perspective of achieving a more solid and wide-ranging evidence of
NBS impacts through on-going and future efforts in collaboratively and
collectively preserving, organizing and sharing NBS related data (Hampton et al.,
2013; see also Section 7.10.4). Examples on the use Big Data for NBS
assessment are provided in, e.g., Section 7.8.

288
7.2 Environmental data of relevance for NBS monitoring and
assessment

In Section 7.1.7, observational data were differentiated into environmental and


population observations and a brief definition of both was provided. This section
focuses on environmental data. A wide variety of approaches has been developed
to observe environmental and ecological impact of NBS, taking experience from
the previous background of the research community in these fields (Houghton et
al., 2012; Lein, 2012). In fact, this represents one of the most established areas
of nature-based solution evaluation.

A diversity of methods has been implemented that cover a broad range of the
potential benefits, and trade-offs, associated with nature-based solution
implementation. In terms of data types, there are two categories of
environmental observations which are essential and widely used to assess and
monitor the physical or environmental conditions of a NBS site and to establish a
baseline: remote sensed data and in-situ observations and measurements. In
some cases, these observations are also complemented by survey data gathered
at the NBS site or available from national databases.

These measurement techniques allow to gather a large variety of environmental


data. In that respect, the concept of “essential climate variables” (ECVs), might
be useful (WMO, 2020). The concept of Climate Essential Variables was first used
for the development of the Global
Climate Observing System. The
essential climate variables (ECVs)
are formally defined as “physical,
chemical, or biological variables or a
group of linked variables that
critically contributes to the
characterization of Earth’s climate”
(Bojinski, 2014). The concept of
essential variable has expanded also
to other domains like biodiversity,
ocean, social sciences, thus

providing an excellent basis for building a NBS monitoring system. Another


advantage of using the ECVs is that currently a lot of research is focusing to
anchor the ECVs to Sustainable Development Goals and other international
initiatives and their targets (e.g., ICCP, Sendai). Some studies place the ECVs
between basic observations and indicators as single EV capturing a key process
or structure can potentially contribute to multiple indicators, while similarly two
or more ECVs can direct and use the same primary observations (Reyers, 2017).

289
Figure 7-2. How can we generate and collect data to evaluate environmental and ecological
impacts of NBS?

7.2.1 Remote sensing (RS) and Earth Observation (EO)

Remote sensing (RS) is the technique of observing and collecting information


about an object or phenomenon from a distance, by means of sensors that are
not in physical contact with the object of investigation (target). The platform
employed to be “at a distance” from the target can be air-borne, space-borne or
ground-based. Typical airborne platforms are drones and aircrafts, while satellites
are used as space-borne platforms. Note that when the target of investigation is
the Earth, the term Earth Observations (EO) is commonly used to indicate data
gathered from Earth observing satellites. Finally, ground- (or sea-) based
platforms consist of sensors mounted on tripods or moving vehicles. These
platforms, along with drones, are mainly used for acquiring very detailed
information at smaller spatial and temporal scales.

At present, a multitude of RS techniques is available, including visible and infrared


imaging, light detection and ranging (LiDAR), and synthetic aperture radar (SAR).
Multispectral sensors allow to study the changes of vegetation or built areas (land

290
use changes), while thermal imagery can be used for measuring the urban heat
island effect. Beside satellite imagery, aerial photography is another important
source of information about the Earth surface: LiDAR sensors, for example, allow
gathering high-resolution elevation data, which can be applied for measuring the
heights of trees or buildings.

Remote sensing and EO are also frequently used to analyse forest dynamics,
pollution level, changes in soil erosion, an estimate of the animal population, and
the impact of natural disasters. In the context of NBS monitoring, they provide
affordable, high quality mapping and monitoring of urban and environmental
parameters at multiple spatial scales (Kabisch et al., 2016). Table 7-1 provides
some key examples of how global Earth observation data can be integrated into
NBS models. It highlights how RS data can be used to improve the understanding
of the processes controlling spatial and temporal dynamics of NBS.

One of the main advantages of RS


and EO is their low-cost and vast
availability which can greatly
contribute to monitoring of NBS. A
list of sources for EO data is
presented in Table 7-3. In general,
freely accessible data (free of cost)
is provided by public agencies,
under potential conditions linked to
the application envisaged and the
nationality of the entity requiring
access. European Space Agency
(ESA) provides detailed
information on access to Earth
Observation data products at
https://earth.esa.int/web/guest/data-access, where products can be browsed by
mission and instrument, or by Earth topic, typology, and processing level. Data
can also be bought from private companies operating commercial satellites, or
by their numerous certified resellers. A unique source of freely available satellite
data is the European Copernicus Program (https://www.copernicus.eu): the vast
amount of EO data relevant to NBS monitoring is divided into 6 thematic domains.
These data are freely available to all users via different channels. One of them is
The Copernicus Open Access Hub which provides free and open access to not only
raw and processed data, but also computational algorithms and cloud computing
facility.

291
Table 7-1. Key examples of how global and European Earth observation data can be integrated into NBS models and how remote sensing can improve the
understanding of the processes controlling spatial and temporal dynamics of NBS.

Theme analysed Which particular data can be Remote sensing data sources Data Provider (SRS data product)
provided

Climate change Contemporary observations of Some satellite remote sensing Gas concentration: Terra/Aqua (MODIS),
ecosystem status and trend, missions provide long-term records of Nimbus‐7/Meteor‐3/Earth Probe (Total Ozone
(remote sensing to together with environmental land surface temperature and of Mapping Spectrometer (TOMS) (1978‐2006),
monitor the rate, models, can help to estimate vegetation, from which indices useful Sentinel‐5P (TROPOMI)
magnitude, and the ecological and economic for understanding the dynamics of
spatial and effects of climate change and climate change can be derived.
temporal effects of to develop and assess
climate on See also: Copernicus Open Access Hub
adaptation and mitigation (Table 7-3)
ecosystems) plans

Data on other high-priority variables, a) Thermal remote sensing, VIs, climate


such as: a) evapotranspiration and b) data; b) RADAR, HSI.
soil texture, moisture and chemistry
are also measured by remote sensing See also: Copernicus Open Access Hub (
Table 7-3)

Time-series data on vegetation a) Biomass, C storage - LiDAR, RADAR,


derived from multiple sensors multiangle RS; b) Photosynthesis, C
contribute to understanding the sequestration - fPAR, photosynthetic
temporal variability and trends in efficiency, fluorescence, MODIS NPP
vegetation processes and their relation
to climate. See also: Copernicus Open Access Hub (
Table 7-3)

Climatological, meteorological, ECMWF


hydrological datasets. Operational,
real-time and re-analysis datasets. Climate Data Store

292
Ecosystem Cost-effective information on Landsat-derived maps for ecosystem Barrier effect of vegetation (forest cover) -
processes ecosystem extent, status, services provision or a potential loss of Landsat (TM, ETM+, OLI) Global forest cover
trends, and responses to ecosystem function. change (200-2012); tree cover - Landsat
(how remotely stressors over large areas (TM, ETM+, OLI) Landsat Tree Cover
sensed ecosystem (e.g. for quantifying Continuous Fields (2000 and 2005)
variables can be ecosystem services inputs and
used to associations between
understand, productivity, nutrient High spatial resolution and frequent Biological control - changes in maximum
monitor, and retention, health benefits etc.) revisits are most useful for NDVI (Terra/Aqua MODIS); pollination
predict ecosystem documenting long-term effects of (vegetation phenology) - Terra/Aqua
response and extreme events, such as severe (MODIS) NDVI; Primary productivity -
resilience to storms, on ecosystem structure, Terra/Aqua MODIS).
multiple stressors) function, and productivity, but
increased spatial and temporal
resolution imagery would likely result
in a finer scale understanding of
ecosystem responses to these events.

Ecosystem To document, monitor, and Regular monitoring of ecosystem a) (AVHRR), Terra/Aqua (MODIS), TRMM
services ultimately predict the extent services such as: a) emissions of (CERES), NOAA AOML Surface CO2 Flux
and condition of certain gases and carbon sequestration and maps (1982–2009), LiDAR, RADAR,
(how remote ecosystem services (e.g. air storage; b) provision of shade and multiangle RS; b) Terra/Aqua (MODIS) -
sensing-derived purification, flood mitigation, shelter: tree cover and plant canopy; MODIS Vegetation Continuous Fields (2000‐
products can be water management, etc.) c) temperature regulation (land and 2013), Landsat (TM, ETM+, OLI) - Landsat
used to value and within a given area under sea surface temperature); d) Tree Cover Continuous Fields (2000 and
monitor changes in current conditions and future precipitation regulation (rainfall, 2005); c) Terra/Aqua (MODIS) - MODIS Land
ecosystem policy scenarios. evapotranspiration); e) water Surface Temperature and Emissivity,
services) regulation: f) Inland water dynamic - Sentinel 3 (SLSTR) for Land Surface
Also, to establish through Change in water stage and water Temperature; d) TRMM (PR, TMI, VIRS,
analysis of remotely sensed body distribution; g) food - CERES) precipitation estimates (1998-2015),
vegetation cover the baselines production of vegetal biomass; h) Terra/Aqua (MODIS precipitation); e)
for provisioning regulatory and food - vegetation indices; provision of Sentinel 3 (SRAL) altimetry; f) Terra/Aqua
cultural services in schemes of clean water, sustainable fisheries, and (MODIS) water mask, Landsat (TM, ETM+) –
payments for ecosystem agricultural productivity with remote global surface water; g) Terra/Aqua (MODIS)
services. sensing from different sources. – net primary production; h) Terra/Aqua

293
(MODIS) - MODIS FAPAR, MODIS LAI,
MODIS Chlorophyll α

Changes in land The global coverage and the Images with high temporal and low MODIS, Landsat, or their combination.
use and land spatial and temporal resolution spatial resolution, such as those from
cover of satellite observations allow MODIS, as well as images with high See also CORINE Landcover (Table 7-2)
mapping of these small- to spatial and low temporal resolution,
large-scale changes. such as those from Landsat, or their
combination.
Note that more nations are
launching satellites with high
spatial resolution (30 m), but Images with high temporal resolution Daily for MODIS and visible infrared imaging
it is still a challenge to (daily for MODIS and visible infrared radiometer suite vs. bimonthly for Landsat
coordinate and calibrate the imaging radiometer suite vs.
imagery from these systems bimonthly for Landsat) capture the
to increase the frequency of timing of vegetation changes, such as
observations. changes in phenology, and changes in
chlorophyll levels.

Species Data on extrinsic Many of these variables are derived Change in biomass, plant traits, land cover
distributions, environmental drivers such as from existing multispectral sensors (Multitemporal RS)
abundances, and land cover, primary (e.g., MODIS).
life stages productivity, density of
human-made structures,
habitat quality for given However, macroscale analysis may a) Species map: Chemical or structural
species. require deployment of new sensors uniqueness, HSI, LiDAR, image texture; b)
such as satellite-based light detection plant traits: spectral analysis or radiative
and ranging (lidar) or 3-dimensional transfer models; c) Spectral diversity of
surface mapping and imaging species (Range or variability of biochemistry,
spectrometers for better NDVI, or reflectance in set of pixels); d)
discrimination of features of Abundance of functional components
heterogeneous terrestrial ecosystems. (Spectral unmixing, MODIS Continuous
Derivation of data at finer spatial and Fields)
thematic resolutions may require
combination with on-site observation

294
Degradation and To detect many types of Landsat data. 1) Fire occurrence and extent: Terra/Aqua
disturbance disturbance that manifest in (MODIS FIRMS), MODIS Burned Area
regimes changes in land cover, air Note that although global availability Product, SPOT VGT Burned Area; 2) flood
pollution, and different effects of hyperspectral data is limited, much occurrence: Terra/Aqua (MODIS) - NRT
of global climate change. progress has been made in the use of Global Flood Mapping, TRMM (CERES) -
hyperspectral data to assess changes Global Flood Monitoring System, DMSP
in ecosystems and function. (SSM/I), ERS‐1, POES (AVHRR) -global
Multi-sensor approaches may be inundation extent from multi-satellites
particularly useful for assessing (1993-2007; 3) drought occurrence: TRMM
changes in ecosystems, especially (PR, TMI, VIRS, CERES) - Satellite‐Based
when combined with ancillary data Global Drought Climate Data Record,
such as field observations and Eutrophication of water bodies - ENVISAT
topographic data. (MERIS), Terra/Aqua (MODIS), Sentinel 3
(OLCI)

See also: EEA Air Pollution Index


(https://www.eea.europa.eu/themes/air/air-
quality-index)

295
This and other available RS and EO data repositories represent a valuable tool for
NBS evaluation, as they offer continuous long-term monitoring, and allow going
back in time (thanks to archived images) and construct a baseline. Furthermore,
thanks to latest technological improvements, high spatial and temporal resolution
and improved accuracy of data can be achieved in some cases. In general, the
following, generally accepted characterization of spatial resolution can be used
for terrestrial applications:

• Low or coarse resolution, >1 km (e.g., advanced very high-resolution


radiometer [AVHRR]);

• Moderate resolution, 250 m–1 km (e.g., moderate resolution imaging


spectroradiometer [MODIS]);

• High resolution, 30 m (e.g., Landsat);

• Very high, approximately a few meters (e.g., IKONOS, Quickbird, and


airborne remote sensing campaigns).

Table 7-3. Earth Observation data sources and their accessibility - selection of representative EO images
providers (source: ESA, 2019).

Satellite data platform Source of EO data Public access / Commercial


providers free of cost data data

ESA https://earth.esa.int v
/web/guest/home

EU Copernicus /Sentinel https://sentinel.esa.i v


operated by ESA nt/web/sentinel/

Sentinel Hub https://www.sentinel v


-hub.com/

EU Copernicus Open https://scihub.coper v


Access Hub nicus.eu/

EU Copernicus Data and https://scihub.coper v


Information Access nicus.eu/twiki/do/vie
Services (DIAS) w/SciHubWebPortal/
WebHome#dias-box

Copernicus portal https://www.coperni v


cus.eu

Eumetsat http://www.eumetsa v
t.int/website/home/i
ndex.html

296
European Environment https://www.eea.eur v
agency (EEA) opa.eu/data-and-
maps

USGS (Landsat) http://earthexplorer. v


usgs.gov/

NOAA http://www.ospo.no v
aa.gov/

NASA https://earthdata.na v v
sa.gov/earth-
observation-data

Digital Globe resellers http://www.digitalgl v


obe.com/partners/ce
rtified-resellers

Airbus https://www.intellige v
nce-
airbusds.com/access
-to-our-products/

Deimos https://www.deimos v
-
imaging.com/imager
y-store/

Planet Labs https://www.planet.c v


om

ImageSat International http://www.imagesa v


NV tintl.com/about-us/

Urthecast https://www.urtheca v
st.com

MDA Geospatial Services http://gs.mdacorpor v


ation.com/Partners/P
artners.aspx

E-geos http://www.e- v
geos.it/index.html

Satellite Imaging http://www.satimagi v


Corporation ngcorp.com/

CGG http://www.cgg.com v
/default.aspx?cid=74
50

European Space Imaging http://www.euspacei v


maging.com/

297
Land info http://www.landinfo. v
com/

Terra server http://www.terraser v


ver.com/

Apollo Mapping https://apollomappin v


g.com/

It is, however, important to notice that satellite observations have constrains and
therefore should be ideally complemented by ground measurements and other
high-resolution RS platforms such as drones. One of the main constraints of
satellite images concerns the shadows due to the size of the frame, which can
hide certain elements of the image and thus generate errors. This is particularly
critical in dense environments such as cities. The drone technology is a viable
way to provide the missing information and overcome this problem, as it offers
the possibility to do 3D reconstruction and accurate geometric measurements.
Indeed, while satellite imagery enables large spatial coverage with sometimes a
resolution too low for the neighbourhood scale, drone imagery will collect high
accuracy data in a more restricted area with the possibility of capturing different
parameters depending on the drone equipment. This is particularly advantageous
when there is a need for very detailed (or specific) and up-to-date information
about the NBS intervention area.

Despite providing unique viewing angles otherwise not possible from manned
aircraft, and representing a highly deployable technology already adopted in
many applications (for humanitarian, safety, and economic reasons or simply for
surveillance, precision agriculture and data/map acquisition), the use of drones
for NBS monitoring remains at present quite unexplored. This is due to several
limiting factors such as citizens safety, data and privacy topics, and the fact that
some types of drone equipment are rather expensive and/or are restricted in
flight limit zones where flight permission are required. Ground measurements, on
the other hand, represent a more common and widely employed option to
complement satellite data and they are also required for the validation of remote-
sensed data. They are inevitable during the full process of NBS development. For
example, to acquire a full cognition of the intervention area, the survey of its
current biodiversity or the built surroundings can be performed only with ground
measurements. This will be further discussed in Sections 7.2.2–7.2.3.

7.2.2 In-situ observations and ground measurements

In-situ (or local) observations is the technique of observing and collecting


information about an object or phenomenon which is in close proximity to the
observer or the measuring device (sensor). When in-situ observations are
acquired by means of sensors placed either on or near the ground (or into deeper
layers of it), then they are usually referred to as ground measurements. Data
acquired through a standard weather station are an example of ground

298
measurements. Weather and other types of field monitoring stations usually
capture a multitude of qualitative and quantitative environmental data on a
continuous basis, including meteorological, hydrological, and chemical
parameters. This approach has the advantage that data are typically collected
using verified scientific methods and can be fed into data modelling processes to
enhance the predictive quality of the data.

In-situ observations and ground measurements can be utilized for the


assessment and monitoring of the surface and subsurface including terrestrial
ecosystems (e.g., biota and soils), assessment of contaminated land, the follow-
up of in-situ remediation technologies (in particular those for soils, vegetation,
groundwater), as well as for monitoring micro-climate variations and air quality
at the NBS site (Gruiz et al. 2017). Relevant data sources of in-situ observations
are given in Table 7-4. These data are generated through dedicated observation
networks which provide long-term and continuous monitoring of various
environmental and physical parameters.

In addition, the recent advancements in smart, low-cost sensors and wireless


technology is allowing to develop dense and low-cost wireless sensor networks in
cities. The Wireless Sensor Networks of Heraklion, Greece, is an example of it 58.
In general, Wireless Sensor Networks (WSN) can be used to measure air
pollution, traffic, meteorological parameters, noise, water quality, animal
tracking, different risks (landslides, forest fires, flooding, earthquakes), impact
of industry (waste monitoring, machine conditions), health conditions (physical
state tracking, health diagnosis). These data can be used as baselines for
evaluation of NBS environmental impacts.

58
http://www.rslab.gr/downloads_urbanfluxes.html

299
Table 7-4. Available data sources for in-situ observations and ground measurements (selection of six
representative observation networks for environmental monitoring).

Name Web link Description

ICOS http://www.icos- Measurement network dedicated to the


(Integrated infrastructure.eu/ monitoring of greenhouse gases budgets in
Carbon 12 European countries since 2008 (Ciais et
Observation al., 2014).
System)

GLEON https://gleon.org/ Grassroots network of limnologists,


(Global Lake ecologists, information technology experts,
Ecological and engineers who have a common goal of
Observatory building a scalable, persistent network of
Network) lake ecology observatories (Weathers et al.,
2013).

FLUXNET https://fluxnet.org/ A global portal which hosts harmonized and


integrated fluxes measurements (ecosystem
carbon, water, and energy fluxes) provided
by more than 800 sites (active or historic)
around the globe. It includes smaller
networks targeting specific land use types,
such as urban area or inland water systems.
Besides fluxes, ancillary atmospheric state
variables, like temperature, humidity, wind
speed, rainfall, and atmospheric carbon
dioxide are also measured (Pastorello et al.,
2017).

European http://www.europe- The database hosts data acquired since 1996


Eddy Fluxes fluxdata.eu/home in the context of previous and ended
Database research projects, mainly funded by EU.
Cluster Datasets include fluxes of different Green
House Gases and ancillary atmospheric state
variables, like temperature, humidity, wind
speed, rainfall, etc.

European https://www.eea.europa. The European Environment Agency gathers


Environment eu/data-and-maps data and information on a wide range of
agency (EEA) topics related to the environment (pollution,
water, climate, etc.)

EC Joint https://data.jrc.ec.europ This catalogue contains a wide range of


Research a.eu/ datasets of all science areas of the JRC
Centre (JRC)
Data Hub

Given the extensive variety of parameters which can be measured through in-
situ observations and the likewise wide range of NBS KPIs (see Chapter 4) which
can be derived based on this data category, it would have been impossible to
provides an exhaustive overview in the context of this handbook. However, it is
important to notice that generation of in-situ observation data can represent a
nature-based solution metric on its own (i.e., quantifying a change in air pollution

300
level by direct measurement) as highlighted by the key examples reported in
Table 7-5. Furthermore, these environmental and ecological data (Table 7-5) are
usually combined together with other measured parameters to create a combined
metric (i.e., making ground observations of tree species, size, and Leaf Area
Index (LAI) to support modelling of air pollution fluxes). However, due to the
scale of the research field related to ground observations and nature-based
solution evaluation, identifying the most appropriate/effective metric can be
challenging. This is where detailed consideration of the NBS type, and associated
theory of change (Chapter 2) are critical.

Table 7-5. Examples of indicators that have the potential to generate data using ground observations and
how they have been used to assess NBS impacts with respect to Challenges 1 (Climate Resilience), 2 (Water
Management), 3 (Natural and Climate Hazards), 4 (Green Space Management), and 6 (Air Quality).

Essential Variables Indicators Challenges


evaluated through
ground
measurements

Direct Measurement Heatwave incidence expressed as the number of 1


of air temperature combined tropical nights (>20°C) and hot days
(>35°C) per annum

Direct measurement • Surface runoff in relation to precipitation 2, 3


of precipitation quantity
volumes and • Flood peak height (m)
stormwater
flowrates entering
and leaving a NBS

Direct measurement Water quality: total metals abatement (% reduction 2


of water quality in metal pollutants with individual metal/metalloid
parameters pollutants selected based on initial conditions)

Direct measurement Number of days during which ambient air pollution 6


of air pollution concentrations in the proximity of the NBS (PM2.5,
parameters PM10, O3, NO2, SO2, CO and/or PAHs expressed as
concentration of benzo[a]pyrene) exceeded
threshold values during the preceding 12 months

Direct measurement Total O3, SO2, NO2, CO removed by NBS vegetation 1, 6


of wind (unit of mass/year): modelled or measured
direction/speed

Direct measurement • Total carbon removed or stored in vegetation 1, 4


of soil quality and soil per unit area per unit time
• Soil organic matter content (%)

Direct measurement • Total carbon removed or stored in vegetation 1, 6


of Tree size and Leaf and soil per unit area per unit time
Area Index (LAI) • Total PM10 and PM2.5 removed by NBS
vegetation (g/m2 per year)

301
7.2.3 Surveys

Surveys are another valuable method of collecting in-situ data relevant to NBS
environmental monitoring. Data acquisition is done through manual sampling
(removal of the soil, water, vegetation, etc.) and samples are then analysed in
laboratories or more often on-site by portable devices or in mobile laboratories
(Gruiz et al., 2017). However, the data are usually accompanied by uncertainties
due to spatial and temporal (in particular, seasonal) heterogeneities typical for
different environmental parameters.

Surveys are essential for studying diverse ecological phenomena (e.g., plant
successions, species’ population dynamics in an ecosystem, lake eutrophication,
etc.) connected with the implemented NBS (Clobert et al., 2018). As an example,
surveys can be used to assess the role of NBS in biodiversity enhancement by
monitoring the abundance of living species in the NBS area and in its proximity.
Indeed, NBS may contribute to enhancing connectivity by creating ecological
corridors in urban context, thus enhancing biodiversity (including rare and
threatened species; Bonelli, 2018; Nieto et al., 2014). Several biodiversity
monitoring protocols have been developed and tested so far, and they are often
adapted to the local needs, based on the NBS type, size, and on the stakeholders
involved. All the reported protocols commonly shared the systematic approach.
Examples of adopted protocols are reported in Table 7-6.

Table 7-6. Examples of biodiversity monitoring protocols (based on the monitoring activities conducted in
the EU-H2020 project proGIreg). Source: Baldacchini (2019).

Category Monitoring Protocols Adopted

Pollinator biodiversity Site: Urban park (Turin, Italy).


monitoring Data sampling is conducted along specific
transects, which allows the recording of
Pollinators play a key role in associations between flowers and pollinators.
every terrestrial ecosystem. They Transect walks also offer the possibility to evaluate the
are pivotal not only from a success of NBS implemented by combining butterfly
biodiversity conservation point of and bee responses at community level. Surveys are
view, but also for food production made from April to September. Windy and rainy days
and for global economy. are avoided for all observations and samplings.
Monitoring this insect group is
very useful to evaluate the Bee surveys: Each survey comprises 250m long linear
environmental status (EU transects walked in 50 min. Each transect start point
Pollinators Initiative 2017, and direction walked were randomly determined. All
Underwood 2017). bees unambiguously identifiable are recorded and all
others are caught for later identification. Bee richness
and abundance are determined. The honeybee is
identified to species level (Apis mellifera) while other
bees are identified to genus level. Surveys are made at
least one per month, between 9:00 am and 5:00 pm.
Flower surveys: Larval food plants and adult nectar
sources of butterflies as well as flower surveys are
carried out in parallel to the bee and butterfly surveys
along the transects.

302
Butterfly surveys: Transects are 300-500 m long,
depending on the investigated area (according to the
“Pollard walk” (Pollard and Yates 1993). Butterfly
species are identified, and individuals of each species
counted. Surveys are made, every two weeks,
between 10:00 am and 3:00 pm for butterflies.

Phytoplankton biodiversity Site: Renatured lake (Ningbo, China).


monitoring Water samples are collected once a week, for
two years, at 3 sampling points, set at the inlet,
Plankton plays an important role outlet and centre of the lake. Samples are
in fisheries, water pollution analysed under the microscope to identify the species
prevention and environmental and number of phytoplankton and zooplankton
impacts of water conservancy individuals present in each sample.
projects (Sun et al. 2018)

7.3 Socio-economic, demographic and behavioural datasets for NBS


monitoring and assessment: Methods and sources

Socio-economic, demographic, and behavioural data are essential in any NBS


monitoring protocol as they allow assessment of the socio-economic and socio-
cultural impacts of NBS, while also offering insight on public perception, degree
of acceptance and aesthetic and/or recreational merit. In the EKLIPSE Working
Group impact evaluation framework, for example, they are required for
evaluation of many KPIs related to Challenges 6-10 59.

A valuable source of data which fall in this category is the Statistical Office of the
European Union, Eurostat (Table 7-7). More generally, these data are usually
available from government agencies such as National Bureaus (or Offices) of
Statistics. However, data retrieved from the aforementioned sources have often
constrains and limitations due to the
unavailability of updated statistics,
especially in small areas such as
neighbourhoods and suburban
areas, or due to the lack of analysis
which target specific data needs for
the implementation and monitoring
of a NBS (e.g., distribution of people
for single age group in small areas).

59
6: Urban Regeneration; 7: Participatory Planning and Governance; 8: Social Justice and Social Cohesion;
9: Public Health and Well‐being; 10: Potential for Economic Opportunities and Green Jobs (see Chapter
5)

303
Table 7-7. Relevant databases of statistical data (incl. socio-economic and demographic)

Name Web link Description

Eurostat https://ec.europa.eu/eu Supplier of a broad range of socio-


(Statistical Office rostat/data/database economic data. Specific data themes
of the European includes economy and finance;
Union) population and social conditions;
industry, trade and services; among
others (full list available at:
https://ec.europa.eu/eurostat/data/brow
se-statistics-by-theme). It also provides
statistics in alignment with the targets of
the Sustainable Development Goals.

Socioeconomic https://sedac.ciesin.colu It includes various types of statistical


Data and mbia.edu/ data (in form of spatial dataset and
Applications Center maps) at global scale, including
(sedac) population density and distribution,
anthropogenic biomes, population
dynamics (migration, fertility, and
mortality), poverty, etc.

OECD Datasets https://data.oecd.org Comparisons by topic and country of


several categories of data

World Bank Open https://data.worldbank. It includes a variety of regional or


Data org/ county-level datasets in tabular format,
vector or raster geographical data and
unit-level data from sample surveys and
administrative systems.

Infrastructure for https://inspire.ec.europ The INSPIRE Knowledge Base was


spatial information a.eu developed after the adoption of the
in Europe INSPIRE Directive (2007/2/EC). The
(INSPIRE) Knowledge Base comprises of datasets
Knowledge Base on multiple environmental, demographic
and socio-economic domains.

Risk Data Hub https://drmkc.jrc.ec.eur The risk data hub is an open access
(DRMKC) opa.eu/risk-data-hub platform for risk related geospatial data
in Europe. The data hub encompasses
current and future hazard and exposure
analysis as well as loss and damage data
of historical events. The data is available
on different scales based on the NUTS
classification in Europe.

ClimateAdapt https://climate- ClimateAdapt offers a list of statistical


adapt.eea.europa.eu/#t and spatial indicators on climate change
-database adaptation. In addition, the database
includes other data types (videos,
publications, case studies and more).

304
In cases when data are unavailable (or inadequate), a customized data collection
is required, which becomes the sole solution for monitoring the socio-economic
performance of the NBS interventions. Under this perspective, a wide range of
collection data methods exists including qualitative analysis (focus group,
observational methods), surveys, and co-participation methods. Therefore, the
following sections encompass the main approaches and methods adopted in this
context and present practical examples of their applications. Although each data
collection method is presented here as standalone, it is important to recognise
that socio-demographic and behavioural data are often and preferably the result
of mixed and integrated approaches which rely on multiple data types and
methods discussed hereafter.

7.3.1 Quantitative, qualitative and map-based surveys

Surveys represent a well-known and widely adopted method of collecting


sociodemographic, economic, and behavioural data. They can be differentiated
into quantitative, qualitative, and spatially anchored (map-based) surveys,
depending on the specific data needs and research approach adopted.

Quantitative surveys are primarily conducted with questionnaires. Following


the definition of Creswell (1999), quantitative research aims at “explaining
phenomena by collecting numerical data that are analysed using mathematically
based methods (in particular statistics)”. Data gathered through quantitative
surveys are indeed – and by definition – expressed in numerical format and
therefore they can be managed and analysed statistically (as opposed to
qualitative survey data which are usually non numeric). The quality of collected
data represents a crucial aspect in a quantitative survey. To ensure quality,
relevance, simplicity, accuracy and clarity of the questionnaire (or any other
measuring instrument) should be carefully verified before the start of the
investigation. Choice of the proper sampling approach (probabilistic vs. not
probabilistic), calibration of the measuring instrument (e.g., questionnaires) as
well as identification of suitable strategies for data collection are also critical
factors to be considered.

Qualitative surveys are primarily conducted with interviews. They are a


common method adopted in qualitative research, which can be described as
explanatory research aiming at understanding a context or underlying reasons
and motivations (e.g. what are people perceiving about an NBS or why are they
perceiving it like this?). In contrast to quantitative data, qualitative data aim at
describing, and not at predicting. They are typically not numerical but can be
analysed using more recent statistical methodologies that do not necessarily
emphasise the numerical aspect but rather the relationships. In general, data
gathered from qualitative surveys are more complex than quantitative ones and
have also constrains in terms of generalisations and upscaling due to the small
size of the population sample investigated. However, tools used for qualitative
surveys are very versatile and have a participatory character. Common tools
include open-end questionnaire, one-person-interview, and focus groups.

305
Focus groups are used to gather a larger number of information emerging from
group discussions on a specific topic and are led by an expert moderator
(facilitator). This measuring instrument has proven to be very useful in the
building-up phases of any process, since it investigates perceptions, opinions,
beliefs and attitude towards a product or process. Although this vast amount of
information is difficult to categorise in a systematic way, it represents a valuable
and effective tool to allow the monitoring and consequent adaptation of NBS
planning and implementation. Focus groups would therefore be a useful
opportunity for enabling people in participating in a real co-design NBS process
and for preventing marginalization and social exclusion in the social-ecological
context in which they are embedded. Furthermore, engaging stakeholder in the
process of decision-making on NBS can, simultaneously, increase the
performance of an intervention (Woroniecki, 2019).

Map-based surveys are online questionnaires that are increasingly used to


enhance public participation as well as co-creation (Linden and Sheehy, 2004).
This type of survey data allows for automatized spatial anchoring of the collected
survey data. It is a participatory tool for collecting primarily socio-economic data
but also for establishing the opportunity for citizens to actively engage in
decision-making and, simultaneously, enhancing transparency, trust and
satisfaction in planning processes. The added value of collecting spatial anchored
survey data for NBS monitoring and assessment can be further highlighted by
considering, for instance, monitoring small scale changes or understanding risk
perceptions which are often place-based. These survey studies can also be
conducted with the aid of various software products currently on the market. An
example is Maptionnaire (https://maptionnaire.com), which is a software for
map-based questionnaire to facilitate public participation. It can be used, for
instance, to learn more about public perceptions and acceptance of NBS. The
software offers a working space for direct data analysis and management.
Furthermore, the data can be exported to shapefiles, XLSX and other data
formats. Another example of a map-based surveys is using crowdsourcing
application Ushahidi (https://www.ushahidi.com/).The application has been
customized within the EU-H2020 project Operandum to collect information about
the exiting NBS installation at the global scale using simple questionnaire with
mapping application (see Figure 7-4 in Section 7.7).

Overall, surveys represent an effective method for collecting both qualitative and
quantitative data relevant for monitoring the sociodemographic, economic, and
socio-cultural system context in which NBS are embedded. Results derived from
the EU-H2020 CONNECTING Nature project offer a meaningful example on how
survey data can be used to assess socio-economic benefits from NBS.
Specifically, the concept of semi-structured interviews using questionnaire was
developed as part of the research work in the project. Data gathered from these
interviews represent an example of ‘process indicators’ since they enable
evaluating the processes involved in successful (and unsuccessful) nature-based
solution delivery.

Figure 7-2 summarises the CONNECTING Nature study and shows the interview
template developed for that purpose. In other cases, such as the EU-H2020-
project Nature4Cities, specific questionnaires are developed in local language to

306
clarify whether the local stakeholders in the pilot cities of the project understand
the benefits and trade-offs of an NBS implementation case.

1. Introduction
•Short history about what this institution and the role of the expert there;
•How do you understand the NBS term? which NBS experiments do you know and which NBS we will discuss?
•How can you classify this NBS (e.g. single case studies, chance examples, on-going labs etc.)?
•What do you consider the most interesting, innovative and transformative case of NBS experiment (e.g. tools,
methods, framework etc.)?
•What do you consider to be the key to success in this experiment? What are obstacles?

2. Description of NBS experiment(s)


•What is the location of the emerging NBS experiment(s)?
•What is/was your role and responsibility in the NBS experiment?
•What actors / stakeholders were involved in the experiment? (Initiating actors, partners, supporters, etc.)

3. Objectives and drivers


•What problem/s did the NBS try to solve? / What need did it respond to?
•What are the most important drivers of the NBS experiment(s)?
•What other categories of challenges does this NBS relate to (e.g. public health and well-being, economic
development potential, green opportunities etc.).

4. Achievements / Multiple benefits / Impacts


•What do you think are the most interesting short-term outcomes / results of this experiment?
•what do you think are the long-term benefits?
•What benefits do you think the experiment had on (e.g. climate change, sustainable development, restoration
of ecosystems and their functions, social cohesion and social integration... or other additional benefits)?
•How these benefits are / were identified and are they being monitored and/or evaluated?

5. Innovative and Transformational aspects of NBS experiment(s)


•What was innovative about the financing of the NBS? What sources of financing were used?
•What was the way of financing the NBS? Are there any financial construction, development plan or scheme?
•Were there any new business opportunities or (green) jobs created as a direct or indirect result of the project?
•What do you think is socially and organizationally innovative about the process of setting up the experiment?
•Did the NBS experiment(s) enhance stakeholder participation and include new (social) learning processes?
•Did the NBS experiment(s) include new types of collaborations for example between different societal sectors?
•Did the NBS experiment(s) included informal or formal networks for the organization and/or collaboration?
•Did the NBS experiment(s) include product or service innovation in terms of novel technologies used?
•Maybe the NBS experiment include novel environmental / ecological aspects/insights thet were used?

6. Monitoring, evaluation and final questions


•Do you know how the experiment(s) is going to be evaluated and monitored? If so, can you explain how?
•Are or did you use any novel monitoring and/or evaluation tool, Database, Cloud and/or Geospatial tools used
for monitoring, controlling and communicating the NBS?
•What do you consider the biggest challenges/problems for emergent NBS experiments?
•Are you familiar with any novel, emerging, particularly interesting experiments outside Europe?
•Do you know any expert or organization that you suggest us to contact and why?

Figure 7-3. Interview template including the six-step iteration applied in the survey of Connecting Nature
project (Dushkova and Haase, 2020).

307
In CONNECTING Nature, experts dealing with implementing NBS in particular
cities were interviewed on emergent, innovative, and novel NBS using templates
(questionnaires). The aim was to identify lessons learned that will benefit other
cities and stakeholders who are interested in designing, implementing, and
stewarding NBSs. The interviews were supplemented by site visits and participant
observation including those during open public events, urban festivals, public
lectures, guided excursions, and other events. The interviews allow to analyse
the following aspects important when planning and implementing a specific NBS:

• Factors of success of NBS examples – what in particular has contributed


to the successful existence of selected NBS examples (e.g., by looking at
the history of their creation, their impact, governance models, methods
of implementation, design and maintenance, additional benefits, costs
and financing);

• Impact of NBS examples on the environment, economics, society and


sustainable development of the city, to better face current societal
challenges, especially the consequences of climate change in cities and
urban regions;

• Trade-offs and conflicts around the NBS – identifying the potential


barriers for the implementation of effective and durable NBS (Dushkova
and Haase, 2020)

Besides, a broad category of computer-assisted approaches has become


increasingly popular for conducting survey studies, such as computer-assisted
web interview and computer-assisted self-interviewing, while more traditional
tools such as paper and pen data collection, or questionnaires by post, tend to
be less used. For example, a web-based survey has been developed in the scope
of the EU-H2020 project EdiCitNet in order to collect data on the social, economic
and environmental performance of Edible City Solutions (ECS) 60. The web-survey
adopts a colloquial and friendly language and has been co-developed with local
ECS, building upon three main scientific theories on the emergence and diffusion
of similar initiatives: strategic niche management, grassroots innovations and
fertile soil (Sekulova et al., 2017; Seyfang and Longhurst, 2016; Wolffram,
2018).

7.3.2 Population observations

Although surveys remain the most popular data collection method used in
research with humans, in-situ observations represent another possible – and
usually complementary – approach for collecting sociodemographic and
behavioural data in connection with an implemented NBS. As explained in Section
7.1, observational tools differentiate from surveys for the fact that data are
collected without interacting with the object of the research: human behaviour is
observed from afar, and it is registered, according to specific, validated protocols.

60
ECS are edible nature-based solutions, i.e., NBS related to urban food production, processing and use

308
This type of in-situ observations is particularly useful when trying to gather up-
to-date and detailed data in small areas such as neighbourhoods and suburban
areas. For example, certain types of NBS such as public parks, urban forests, tree
corridors, renatured river or lake shores, have the benefit (or co-benefit) to
provide (or provide access to) a space that the population can use to visit green
and/or blue spaces and/or for physical activity. To evaluate whether this is
effective, systematic observation can be performed on-site in order to monitor
the use of the NBS and to assess the related changes in time (before and after
NBS implementation).

A method to quantify the use of a green/blue space is, for instance, the validated
SOPARC (System for Observing Play and Recreation in Communities) tool
(McKenzie et al. 2006; https://www.rand.org/health-
care/surveys_tools/soparc/user-guide.html). SOPARC can provide data on the
number of users and type of physical activity, which represent a common data
requirement for Challenges 7 (Place Regeneration) and 11 (Health and
Wellbeing), and related indicators.

To summarise the method, trained observers (possibly including participation of


stakeholders) count the number of users at the NBS site and register the users’
characteristics (sex and age group) and type of activity (e.g., sedentary, walking,
or very active). These observations are systematic and periodic; measurements
are taken in specific periods of time (morning, lunchtime, afternoon, and evening)
and specific days (within one week). These periods are defined to get an overall
estimate of the use of the site.

To evaluate the change in use and physical activity, systematic observations can
be performed before and after the NBS implementation is monitored, taking care
of repeating the data collection in the same season. In the case of NBS
implementations in pre-existing public green area, a single post-implementation
SOPARC assessment can be conducted, to describe NBS users and their
behaviour.

While in-situ observations such the one collected with SOPARC provide standard
quantitate data, other methodologies exist which also provide qualitative data.
For examples, methodologies which integrate visual techniques such as
photography, film, video, painting, drawing, collage, sculpture, artwork, graffiti,
advertising, and cartoons are increasingly used in multiple disciplines (Pain,
2012). These methodologies can be used to measure in an indirect way the
crowding of parks without quantitative research: the longitudinal mapping of
graffiti can be considered as proxy of artistic expression or cultural dimension.

7.4 Data sources for the assessment of changes to health and


wellbeing

There is an increasing recognition of NBS co-benefits as influential determinants


of human health and well-being (Barton and Grant, 2006; Hartig et al., 2014;
Kabisch et al., 2017). They relate to the provision and improved availability of
urban green spaces and may result in better mental and physical health. A great

309
number of the scientific literature provides results of how different urban nature-
based solutions can affect the health of urban residents and present
epidemiological evidence of public health benefits of green spaces (Beyer et al.,
2014; ten Brink et al., 2016; Dushkova and Ignatieva, 2020; Frumkin et al.,
2017; Groenewegen et al., 2006, Kabisch et al., 2017; Kabisch and Haase, 2018;
Marcel et al., 2019; Williams, 2017; Wood et al., 2016). There are three urban
health dimensions, namely environmental conditions and related health
outcomes, urban equity and vulnerability as well as resilience to extreme climate
conditions related to climate change.

There are many direct links between nature and human health and well-being
which resulted from the epidemiological surveys. Thus, connection with nature,
in addition to satisfying elementary human needs (e.g., food and natural
resources supply), heals or mitigates the most diseases and can be defined as a
health resource (which keeps people healthy) (Groenewegen et al., 2006;
Kabisch and Haase, 2018). The
recreational and healing value of
nature for physical health and
mental well-being has long been
discussed (Beyer et al., 2014;
Hartig et al., 2014; Marcel et al.,
2019). However, nature also has
another value for health,
regardless of natural remedies
(though often not consciously
perceived). For example, the
healing of space, outdoor training
trails in parks, everyday use of urban green spaces and peri-urban recreation
areas for sport and exercises (cycling, jogging, and Nordic walking). These health
aspects of outdoor nature are used for promotion healthy life-style, especially for
children, through the active nature experience, since many children in urban
spaces no longer have the opportunity to acquire nature in everyday life
experience (Kabisch and Haase, 2018). Thus, as a source of healing, and source
of inspiration, nature plays an important role in the identity of people and in the
development of its own "sense of place" (Frumkin et al., 2017).

While the provision of nature-based solutions refers traditionally to environmental


organizations and planners, greater involvement of the health sector will be
important for maximizing benefits for both health and nature. Integrating policy
on biodiversity, health and urban planning to realize joint benefits requires data
from all fields to be linked and communicated to policy makers, to be considered
in impact assessments and economic valuation of decisions (Kabisch et al., 2017).

Main types of data needed to study the relationship between NBS and human
health are:

• Quantitative data from case studies – epidemiological survey and regional


statistics; often, local practitioners benefit from quantitative data and it
is helpful to consider early in the process what quantitative data could be
obtained with reasonable effort. The use of routinely collected statistical
data on local level should be maximized. Yet, the use of other types of

310
arguments and measurements to complement the quantitative data is
necessary to avoid that the lack of quantitative data is interpreted as a
lack of evidence in general;

• Qualitative data (e.g., from semi-structured interviews) which can allow


to capture all the needs of the varying community subgroups. The
interviewing of the intended users of the intervention could be a good way
to gain understanding of their needs as well as their experience with
similar NBS implemented earlier in another place. Various techniques can
be used to collect these data such as using maps during interviews to gain
a robust understanding on how people use and move in and around local
green space.

Literature review shows that very often the following study design was applied:

• Cross-sectional questionnaire survey of women or/ and men (mostly


separating adults from children). Stratified random or cluster sampling
design;

• Observational study of the usage of urban parks or other NBS - direct


observation of park users as well as interviews with persons;

• Survey data combined with GIS and green space data, and their analysis;

• Ecological study of mortality and dasymetric mapping of air pollution and


greenness;

• Observational ecological study comparing neighbourhood socioeconomic


status of women and individual physical activity;

• Self-administered survey of persons on their perceived general health and


the characteristics of their living environment;

• Health interview survey of persons that examined self-reported health,


social contacts, and characteristics of the respondents' living
environments.

Several guidelines were established by WHO (2017) for simple data collection
methods to identify and assess the value of urban green and other nature-based
solution for human health and well-being:

• Use observational data as a relatively simple and cost-efficient way to


assess how many people are using green space, what types of people are
using it, who they are using it with, for what purposes etc.;

• Use existing audit and observational tools to collect information on play


and recreation in public areas;

• Consider simple and innovative monitoring techniques (e.g., user


satisfaction counters like seen in public facilities);

311
• Engage with local networks and organizations as a way to collect feedback
from community and green space users (e.g., engage with community
councils or committees);

• Collaborate, where possible, with academic institutes and research


centres which can aid with delivering effective monitoring and evaluation
for the intervention as well as cost-efficient monitoring (e.g., through
developing student research projects around the NBS intervention).

It is important to consider existing, routinely collected datasets and how these


might be utilized. Some national or local municipality surveys may already have
baseline information on how people currently use and value local NBS, what
effects were reported and analysed. Good demographic data on local residents
and intended users of the green space is critical for informing the planning and
design of the intervention.

Often, socioeconomic status data but also other data (e.g., on environmental risk
exposure, age and sex, or ethnic and other sociocultural parameters) are
available through standard processes on local level. Such data may often be
available in aggregated form for an urban/neighbourhood area rather than as
individual data. In such cases the smallest-possible spatial unit should be
considered, since understanding the population profile is important to define
equity issues (WHO, 2017).

The role of citizen science and participatory research in evaluation should be


considered. This may aid data collection and evaluation, and would also help to
increase the active uptake of the NBS interventions.

The literature reports on positive health associations for a diverse range of NBS
interventions such as street trees, green space establishment on vacant lots and
greening school playgrounds. Reported benefits in terms of reduced exposure to
air pollution are substantial, and usually complemented by others of social (green
spaces for the public) and/or economic nature (new job and business
opportunities).

However, implementation of urban green infrastructure can result in negative


impacts on the local air quality such as the direct emissions of pollen, fungal
spores and biogenic volatile organic compounds (bVOCs). It is thus of paramount
importance an informed choice of the most appropriate species prior to
deployment. The scale and physical dimensions of the deployment are also critical
and need to be assessed case by case, and the outcomes of similar green
infrastructures may vary considerably in different urban environments (Kumar et
al., 2019).

However, it is important to think in a broader sense when planning NBS


interventions. This means to realize the opportunities for collaboration with
institutes such as schools, universities and health services which may enable
access to relevant data sets and help with informing the design of the
intervention. The potential of NBS co-design activities with schools and
universities has been, for example, demonstrated in one of NBS being
implemented within the framework of the EU-H2020 project OPERANDUM: these

312
activities have shown to have the multiple benefit to introduce climate action in
education with potential positive impacts towards the realisation of the objectives
of SDG11 and SDG13. Also, broader interventions (such as urban extensions,
large infrastructure projects or masterplans for residential areas) could consider
and include urban green space and be informed by the benefits of such provisions.

7.5 Predicting the present and future impacts of NBS with modelling
techniques

Modelling is a critical and often compulsory aspect of NBS impact assessment


(Figure 7-3). It allows to simulate the efficiency of one or more components of
NBS, and to monitor and evaluate progress towards its goals. Here, the term
modelling is employed to denote any type of modelling for any Essential Variable.
Various modelling approaches, from lumped to distributed models, require a
varying level of complexity of the described environment.

Modelling NBS addresses the


representation of processes that
occur in the real world in space and
time. The processes resulted or
caused by NBS transform the
environment through time and can
be mostly described by dynamic
models based on differential
equations. The spatial interactions
of different elements of NBS and
NBS with the environment are
mostly managed by geographic information systems (GIS). GIS can be used to
provide input variables required by simulation models and yield visualization and
analysis of output data. Other ways are represented by direct integration of
numerical modelling which is a mathematical representation of a physical (or
other) behaviour, based on relevant hypothesis and simplifying assumptions.
Various simulation tools together with GIS are used to demonstrate modelling of,
for instance, surface water pollution, spatiotemporal analysis of air pollution data,
modelling of land use changes (Cohen-Shacham et al., 2016). Another type of
modelling – physical modelling – is used to validate numerical modelling data;
the use of physical models supports the understanding design concepts and
processes. Modelling combined with scenarios provides insights into drivers of
change, potential implications of different trajectories, and options for action
(Sang, 2020). Section 7.1.7 presents a more detailed discussion on the modelling
approaches and their complexity.

Modelling approaches are primarily adopted for one or more of the following
purposes:

• Identify and/or understand the underlying processes which


describe with certain level of uncertainty relevant environmental
(or behavioural) response/change of the urban system before
(baseline) and after the NBS intervention. For example, models can

313
simulate different natural processes such as crop growth, flooding, and
local climate regulation (e.g., Mohareb et al., 2012) by green space or
soil nutrient flow. In that respect, the advantage of modelling techniques
relies on the possibility of changing input data and parameters to be in
the model. This allows to understand cause-effect relationships and to
make predictions at a level which is not possible with observations.

• Identify vulnerable urban areas and/or areas which are more


prone to certain natural hazards (e.g., flooding). When implementing
nature-based flood protection, for example, it is essential to conduct a
probabilistic hydrological and hydraulic modelling assessment and map
flood zones with the potential intensity and location of all relevant types
of flooding (Mason et al., 2007; Pregnolato et al., 2016; Raymond et al.,
2017). Such resulted maps of potential inundation will present a range of
return periods and appropriate planning needs. Other techniques include
modelling of flood peak reduction (Iacob et al., 2014) or modelling of
options for stormwater management in the urban environment, including
the quantification of SuDS benefits with the BeST model (Morales-Torres
et al., 2016).

• Generate (or use) simulation data to fulfil the data requirements


for specific KPI, especially when other data collection methods are
not feasible/too expensive, or data are simply not available or
adequate. For example, gross and net carbon sequestration of urban
trees can be estimated with the iTree Eco model (Baró et al., 2014), which
provides a database on ecosystem services rendered by different trees
species in different climatic zones.

• Improve awareness and perception of NBS co-benefits and


efficacy through scenario and impact modelling. For example,
superior performances and co-benefits of a specific NBS versus more
traditional interventions (e.g., grey infrastructure) can be verified through
modelling studies (e.g., Gittleman et al., 2017), and effective
communication of these results may enhance acceptance and
engagement among stakeholders and policymakers. In that regard, it is
worthwhile to mention that models not only represent the environmental
impact of NBS, but they can also model the societal responses and
participatory process by applying methods such as geodesign. As stated
by Steinitz (2016), geodesign helps to find consensus around plans with
sufficient detail to be workable, adaptable to the local needs and context
and sustainable over time. Additionally, development of innovative social
models for long-term positive management (e.g., Citizen Engagement for
Health; Fernandez et al., 2015) may also contribute to increasing
stakeholder awareness and knowledge about NBS and ecosystem
services, as well as citizen participation in the management of NBS
(Filibeck et al., 2016; Hansen et al., 2016).

• Develop design scenarios for the selection of the optimal NBS


among the ones conceivable, and for estimation of efforts needed for its
implementation and maintenance.

314
• Forecast NBS performances and impacts over time and/or in
connection with future climate projections. In this regard, extensive
research modelling efforts have been made to assess effectiveness of NBS
in tackling challenges such as climate change, food security and water
resources. Furthermore, the use of natural hazard modelling has been
expanded and combined with numerical weather prediction and climate
models to develop climate change adaptation and disaster risk reduction
strategies that are resilient, adaptable, resource efficient, locally
adjustable and optimised.

15.

Figure 7-4. Simulation of hydrodynamic and morpho-dynamic processes to assess the effect of NBS
(artificial sand dune) on wave propagation. Left: Numerical Domain showing the position of the NBS along
the shoreline. Right: model results showing the wave propagation (source: EU-H2020 project OPERANDUM;
image credit: ARPAE-IT)

Despite their numerous advantages and countless applications, modelling


techniques have also limitations and uncertainties which should never be
neglected in the evaluation process and/or while using modelling results. Some
of these limitations and uncertainties are intrinsic to the technical or
mathematical structure (or logical framework) on which the model is built, and
on the assumptions and/or approximations which may be embedded into it. For
that, simulation results must be compared to and validated against observational
data to ensure the validity of results and also the quantification of the overall
uncertainty of the simulated scenario assessed. Errors and/or misleading results
can also be generated by an “inappropriate” use of the model. Indeed, every
model is built to address only specific research questions and is meant to be used
only for certain specific applications and contexts. Knowledge of the model goals
and capabilities is thus crucial in order to select “the right tool for the right
problem”.

A variety of numerical models exists that are used to simulate the state variables
such as temperature, precipitation and evapotranspiration. Table 7-8 lists the
relevant modelling tools for assessing ecosystem services provided by NBS. A
non-exhaustive list of the most widely used numerical models can be classified
under the following Challenge areas:

315
• Climate resilience
 General circulation models (GCM) (Mechoso and Arakawa, 2015)
 Weather Research and Forecasting Model (WRF) (Surussavadee et al.,
2017)
 complex numerical methods describe the interactions between
vegetation and pollutants at the micro scale (Joshi and Ghosh, 2014) or
simulate the emission and deposition processes based on trajectory and
dispersion models, e.g. the atmospheric transport FRAME (Fine
Resolution Atmospheric Multi-species Exchange) model (Bealey et al.,
2007).

• Water management
 MIKE11 (Thompson et al., 2017)
 Soil Water Assessment Tool (SWAT; Arnold et al., 2012)
 Storm Water Management Model (SWMM) (Rossman, 2015)
 MODFLOW model (Langevin et al., 2017)
 GREEN (JRC) (Grizzetti et al., 2012)

• Natural and climate hazards


 Discrete Element Method (DEM) (Mahmood and Elektorowicz, 2016)
 ADvanced CIRCulation (ADCIRC) (Luettich et al., 1992)
Table 7-9 presents a selection of studies obtained from the scientific literature,
which show the ways simulation and modelling can be applied to the assessment
of NBS impacts in the urban environment.

316
Table 7-8. Modelling tools for the assessment of the ecosystem services provided by NBS.

Tool/Model Description Source Comment

Artificial A networked software technology that redefines http://aries.in ARIES is meant to enable simple use of complex
Intelligence for ecosystem service assessment and valuation for tegratedmode models through artificial intelligence; as such,
Ecosystem decision-making, to map natural capital, natural lling.org/ extensive training (annual intensive modelling
Services (ARIES) / processes, human beneficiaries, and service flows to schools) is only necessary for modellers who
probabilistic model society as a new way to visualize, value, and manage want to contribute to, and benefit from, ARIES
the ecosystems on which the human economy and models and data.
well-being depend; to quantify the benefits that
nature provides to society

The Atlas of Up-to-date platform for knowledge and information www.atlasnat Companies, governments and citizens can use
Natural Capital dissemination enhancing the sustainable use of uurlijkkapitaal data from ANK
(ANK) / natural capital (currently more than 150 maps on .nl/en
Spreadsheet ecosystem services in the Netherlands)

The Ecosystem A collection of spatially explicit models to support the Zulian et al. It is based on the ecosystem services cascade
Services Mapping mapping and modelling of ecosystem services at (2014) framework which is used as a frame for
tool (ESTIMAP) / European scale. Its main objective is to support EU mapping; it includes four complete models:
GIS application policies with spatial information on where ecosystem outdoor recreation, crop pollination, coastal
services are provided and consumed. protection and air quality regulation.

Benefits Estimation Benefits Estimation Tool – valuing the benefits of https://www. A free tool and guidance for use on PCs. It
Tool (B£ST) / blue-green infrastructure. It assesses and monetizes susdrain.org/r makes assessing the benefits of blue-green
Spreadsheet many of the financial, social and environmental esources/best infrastructure easier, without the need for full
benefits of blue-green infrastructure; it enables users .html scale economic inputs; it can support
to understand and quantify the wider value of investment decisions and help to identify
Sustainable drainage systems and natural flood stakeholders and find potential funding routes.
management measures

i-Tree (formerly Based on peer-reviewed, USDA Forest Service https://www.i i-Tree is a combination of science and free tools;
Urban Forest Research, it offers several desktop and web-based treetools.org/ it provides users/managers with tools by
Effects Model) / applications to quantify the benefits and values of allowing them to improve tree and forest
Desktop software trees around the world, to aid in tree and forest management, plan strategically, increase

317
management and advocacy, to show potential risks awareness, engage decision makers and build
to tree and forest health new partnerships.

ESValues / A collaborative platform that collects economic data https://esvalu It allows users to obtain economic values for the
Spreadsheet from ecosystem services studies to produce value es.org/ ecosystem services provided by an ecosystem
estimates by benefit transfer and upload the parameters and estimates from
these economic valuations

Integrated A suite of models used to map and value the goods https://natur Free, open-source software models; it enables
Valuation of and services from nature that sustain and fulfil alcapitalproje decision makers to assess quantified tradeoffs
Ecosystem human life. It helps explore how changes in ct.stanford.ed associated with alternative management
Services and ecosystems can lead to changes in the flows of many u/invest/ choices and to identify areas where investment
Tradeoffs (InVEST) different benefits to people. The toolset includes in natural capital can enhance human
/ GIS software distinct ecosystem service models designed for development and conservation.
terrestrial, freshwater, marine, and coastal
ecosystems, as well as a number of “helper tools” to
assist with locating and processing input data and
with understanding and visualizing outputs.

A Global Standard Developed by IUCN in order to create a common https://www.i Not yet available (still in the developing stage)
for Nature-based understanding and consensus on Nature-based ucn.org/them
Solutions / Solutions, the Ecosystem Management Programme e/ecosystem-
Spreadsheet and Commission are jointly leading the collaborate management/
process of elaborating a Global Standard for the our-work/a-
Design and Verification of Nature-based Solutions. global-
standard-
nature-based-
solutions

Land Utilisation An ecosystem services modelling tool which https://www.l LUCI is relevant for a range of users at multiple
Capability illustrates the impacts of land use on various ucitools.org/ scales and levels of decision-making. It can be
Indicator (LUCI) ecosystem services. It runs at fine spatial scales and applied for applications around sustainable
compares the current services provided by the development, conservation, sustainable
landscape with estimates of their potential capability. tourism, restoration, and policy-making.
LUCI uses this information to identify areas where
landscape usage change might be beneficial, and

318
where maintenance of the status quo might be
desirable.

The NATURVATION The Naturvation Index (proposed by the EU-H2020 https://naturv Value and Benefit Assessment Methods
index / project NATURVATION) to evaluate nature-based ation.eu/asse Database and Framework for Urban Nature-
Spreadsheet solutions projects and identify how they contribute to ssment based Solutions
sustainability goals.

Social Values for A GIS Application for Assessing, Mapping, and https://solves SolVES derives a quantitative, 10-point, social-
Ecosystem Quantifying the Social Values of Ecosystem Services .cr.usgs.gov/ values metric, the “value index”, from a
Services (SolVES) – SolVES 3.0 tool which is ArcGIS 10-compatible. combination of spatial and nonspatial responses
/ GIS application to public value and preference surveys and
calculates metrics characterizing the underlying
environment, such as average distance to water
and dominant land cover.

The Economics of The Manual presents an overview and explains the http://www.t It allows for user the evaluation of ecosystem
Ecosystems and potential uses and functions of the TEEB Valuation eebweb.org/p services, but not measure the quantities and not
Biodiversity (TEEB) Database. The Manual discusses the origin of the ublication/tth allows to input the data
Valuation Database database; describes its content and structure; e-economics-
/ Spreadsheet outlines its contents and discusses how it may be of-
used, including important caveats. ecosystems-
and-
biodiversity-
valuation-
database-
manual/

Toolkit for The toolkit provides practical guidance on how to http://tessa.t It emphasizes the importance of comparing
Ecosystem Service identify which services, what data are needed to ools/ estimates for alternative states of a site (for
Site-based measure them, what methods or sources can be used example, before and after conversion to
Assessment to obtain the data and how to communicate the agriculture) so that decision-makers can assess
(TESSA) / results. The toolkit has attempted to find a balance the net consequences of such a change, and
Spreadsheet and between simplicity and utility and can be used by hence the benefits for human well-being that
GIS application non-experts, yet still provide scientifically robust may be lost through the change or gained by
information. conservation.

319
Copernicus, Corine Copernicus Land Monitoring Service portfolio (both http://land.co Corine Land Cover (CLC) 2012, Version 18.5.1.
Land Cover by EEA already operational and upcoming) products are pernicus.eu/p Processed by The European Topic Centre on
(European divided into the following categories: an- Land Use and Spatial Information
Environment Land Cover and Land Use Mapping european/cori
Agency) Hot-spot Monitoring ne-land-
Biophysical Parameters cover/clc-
Imagery, In Situ and Reference Data 2012/view
European Ground Motion Service

The assessment of The assessment results helps explaining better to the https://ec.eur The guide has four main components: 1. An
ecosystem and general public and stakeholders the multiple benefits opa.eu/enviro introduction to key concepts and methodology.
their services – of LIFE projects in connection to society and the nment/archiv 2. The description of a simple approach to
approaches from economy with which they interface. The document es/life/toolkit/ assess ecosystem services applicable to all LIFE
LIFE program of clarifies key concepts and offers an easy method to pmtools/life2 projects independently from the method used to
European implement ecosystem services assessments 014_2020/ec quantify them. 3. Guidance on how to complete
Commission according to the analytical framework developed osystem.htm the relevant sections in the LIFE KPI database.
under the EU Mapping and Assessment of Ecosystems 4. A selection of further resources
and their Services (MAES) initiative. Some guidance (https://ec.europa.eu/environment/archives/lif
on how to complete the relevant sections in the KPI e/toolkit/pmtools/life2014_2020/documents/lif
Webtool is also given. e_ecosystem_services_guidance.pdf)

Land Use-based LUISA is developed by Joint Research Centre (JRC) of https://public


Integrated the European Commission, which is primarily used for ations.jrc.ec.e
Sustainability the ex-ante evaluation of EC policies that have a uropa.eu/repo
Assessment’ direct or indirect territorial impact. At its core is a sitory/bitstrea
modelling platform discrete allocation method that allocates different m/JRC94069/l
(LUISA) / GIS land uses to most optimal 100m grid cells, given b-na-27019-
based modelling predefined suitability maps, regional land demands en-n%20.pdf
platform and the supply of land in a region. Linked to the
allocated land uses are grid cell population counts,
which are modelled separately prior to the land-use
allocation. The chief outputs that LUISA generates
are projected land use, population and accessibility
distributions at the 100m grid cell level. Over 50
indicators of land functions are subsequently derived
from those chief outputs. Those indicators can inform

320
policy effects on themes as varied as resource
efficiency, ecosystem services and accessibility.

Integrated system KIP INCA aims to develop the first ecosystem https://public
of Natural Capital accounts at EU level, following the UN System of ations.jrc.ec.e
and Ecosystem Environmental-Economic Accounting- Experimental uropa.eu/repo
Services Ecosystem Accounts (SEEA-EEA). The application of sitory/bitstrea
Accounting (KIP the SEEA-EEA framework is useful to illustrate m/JRC87585/l
INCA) / ecosystem accounts with clear examples and b-na-26474-
Spreadsheet contribute to further develop to methodology and en-n.pdf
give guidance for Natural Capital Accounting.

321
Table 7-9. Studies of the impacts of NBS, which show how numerical simulations and modelling can be applied.

NBS Study Simulation model Findings

Air quality Hirabayashi i-Tree Eco estimates air pollution removal by trees based on well- It allows to quantify the structure of,
et al. established deposition models and hourly air quality and wind threats to, and benefits and values
(2012), speed data from local weather stations provided by forests.
Nowak et al.
(2014) i-Tree Eco: https://www.itreetools.org/tools/i-tree-eco

McDonald et The fine resolution atmospheric multi-pollutant exchange Tree planting was simulated by
al. (2007) (FRAME) atmospheric transport models designed to predict the modifying the land cover database,
impact of NBS implementation of air quality level, e.g. to estimate using GIS techniques and field surveys
deposition of nitrogen, heavy metals and the surface to estimate reasonable planting
concentrations of greenhouse gases by tree planting potentials and predict increasing total
tree cover
FRAME: https://frame-online.eu/

Matos et al. To model the supply of air-quality regulation based on urban A model allows to estimate the supply
(2019) green spaces characteristics and other environmental factors of air quality regulation provided by
(lichen diversity in urban parks) green spaces in all green spaces of
Lisbon based on the response to the
following environmental drivers: the
urban green spaces size and its
vegetation density. The model helps to
map the background air pollution

Bruse Microscale simulations employed for street-scale evaluation with the newest version of the microclimate
(2007), software such as ENVI-MET. model ENVI-met was compared against
Simon et al. measured data
2019 ENVI-MET: https://www.envi-met.com/

Bagheri et FRAGSTATS software (Spatial Pattern Analysis Program for The model results indicate that
al. (2017) Categorical Maps) and a partial least square (PLS) model reduction in the area of large green
space patches promote air pollution,

322
were applied to assess the effects of changes in the pattern of suggesting that there is a direct
green space on air pollution. relation between increases in the area
of large green space patches and air
FRAGSTATS: pollution reduction.
https://www.umass.edu/landeco/research/fragstats/fragstats.html

Book on PLS: https://doi.org/10.1007/978-3-319-64069-3;


https://core.ac.uk/download/pdf/20267242.pdf

Green roofs Bass et al. Use of Mesoscale Community Compressible (MC2) model, A green roof strategy consisting of
for (2002) land use grid cell data, urban canyon model for Toronto, Canada. grass roofs (only 5% of the total city
temperature https://www.coolrooftoolkit.org/wp- area) reduced temperatures by up to
reduction content/uploads/2012/04/finalpaper_bass.pdf 0.5°C. Irrigating green roofs in the
high-density areas produced a much
more intensified cooling effect: 1-2°C
temperature reduction.

Chen et al. Coupled simulations of conduction, radiation and Installing grass roofs on medium and
(2009) convection for Tokyo, Japan high-rise buildings has a negligible
effect on the street level air
temperature.

Smith and Weather research and forecasting model (WRF) coupled with Vegetative rooftops reduce evening
Roebber an urban canopy model (UCM) applied for Chicago, US and night-time temperatures by 3°C
through increased albedo and
(2011) WRF–UCM: https://ral.ucar.edu/solutions/products/urban- evapotranspiration.
canopy-model

Sun et al. Numerical model ENVI-met and verified using field The maximum cooling effect of green
(2012) measurements adapted for Taiwan roofs on ambient air temperature was
1.6°C
ENVI-MET: https://www.envi-met.com/

Urban land Haase et al. Combination of system dynamics (SD), cellular automata Using the example of urban shrinkage,
use (2012) (CA) and agent-based model (ABM) approaches to cover the it highlights the capacity of existing
land-use modelling approaches to

323
main characteristics, processes and patterns of urban land use integrate new social science knowledge
and shrinkage in Leipzig, Germany in terms of land-use, demography and
governance.

Schwartz et It presents the ABMland - a tool for collaborative agent-based ABMland allows for implementing
al. (2012) model development on urban land use change which allows for agent-based models and parallel model
explicitly coding land management decisions. The software is development while simplifying the
implemented in Java building upon Repast Simphony and other coding process. The models include six
libraries. major agent types: residents, planners,
infrastructure providers, businesses,
ABMland: https://www.ufz.de/index.php?en=37897 developers and lobbyists. Their
interactions are pre-defined and ensure
valid communication during the
simulation.

Brown and Rule-based models developed for sectoral strategies such as Such approach of translating scenarios,
Castellazzi woodland expansion, wind energy, urban development as input for storylines and policy objectives into
(2014) development scenarios using LandSFACTS software and the spatially explicit realization can be used
Integrated Agriculture and Control System (IACS) data in a with any spatial unit (land use or cover
stochastic process. polygon, population ward, water
catchment) to explore alternative
LandSFACTS: options for land use and the role of
https://www.hutton.ac.uk/research/departments/information- particular NBS intervention.
and-computational-sciences/tools/landsfacts/downloads

IACS: http://ec.europa.eu/agriculture/direct-
support/iacs/index_en.htm.

Hamad et al. Land use change scenario simulation using a CA-Markov model The models can support to optimize
2018 as one of the commonly used models among many LULC urban land use layout and assist with
modelling tools and techniques decision-making

Water World Bank Modelling NBS for managing freshwater resources Models for provision of safe drinking
management (2017) water, integrated river basin
management, pollution management

324
Brunetti et Surrogate-based modelling for the numerical analysis of low The hydraulic behaviour of the green
al. (2016, impact development techniques roof, permeable pavement and
2017) stormwater filter were analysed by
means of a model approach

Sahukhal The water demand and supply modelling were conducted using The performance of the model was
and the water evaluation and planning (WEAP) model, based on assessed through statistical measures
Bajracharya, discharge data (can be obtained from Department of hydrology of calibration with the root mean
2019 and meteorology). square error and coefficient of
determination. It allows to create
WEAP: https://www.weap21.org/ different scenarios important for the
analysis regarding the prioritization of
demands in the near future for the
purpose of sustainability of water
resources, due to climate change
impacts.

Natural Li et al. The study used the Soil and Water Assessment Tool (SWAT) The SWAT model was forced with
Hazards (2019) module of a GIS platform to simulate the potential of wetlands meteorological variables such as daily
against flood and droughts rainfall, temperature, wind speed,
relative humidity, solar energy and it
SWAT: https://swat.tamu.edu/ was found that restoration and
reconstruction of wetland can reduce
the impact of flooding and hydrological
droughts.

Vuik et al. Modelling the effect of vegetation on flood wave attenuation using The study forced SWAN numerical
(2016) the Simulating WAves Nearshore (SWAN) model wave model with bathymetry, ocean
current, ocean water level, bottom
SWAN: fraction, and wind speed datasets to
http://swanmodel.sourceforge.net/download/download.htm simulate and evaluate the effect of
vegetation on flood wave attenuation.
The datasets were retrieved field
measurements performed on two salt
marshes (cordgrass and grassweed)
during the severe storms in the

325
Netherlands from November to June
2014

Wamsley et Use of the three-dimensional numerical model ADvanced The study simulated the role of
al. (2010) CIRCulation (ADCIRC) to evaluate the role of wetlands in wetlands in reducing storm surges and
reducing storm surges. concluded that wetlands may have
capacity to reduce surges, but their
ADCIRC: https://adcirc.org/ effectiveness depends on the
surrounding, coastal landscape and
the strength and duration of the storm
forcing

Stark et al. Use of the two dimensional hydrodynamic model The study simulated the potential of
(2016) TELEMAC2D to evaluate the role of wetlands during storm tides. wetlands in attenuating peak water
level during storm tides. The result of
TELEMAC2D: simulation showed that peak water
http://www.opentelemac.org/index.php/presentation?id=17 level reduction largely varies among
individual flood events and between
different locations in the marsh, but
the tidal wetlands in combination with
dikes provides more effective coastal
protection

Guida et al. Combination of hydrodynamic (e.g., 1D HEC-RAS) and The study performed two scenarios
(2015) geospatial modelling (e.g., HEC-GeoRAS) to simulate the such as levee removal and leave
optimal flood risk reduction measures for the Lower Tisza River in seatback to reconnect wetland and
Hungary. found that the wetland reduced flood
heights and potential damage to
The main modelling tools and software used in the study are human populations.
developed by the Hydrologic Engineering Center (HEC) - US Army
Corps of Engineers, and available here:
https://www.hec.usace.army.mil/software/

Sang (2020) Integrating Computational and Participatory Scenario Modelling for Comparative review of a wide range of
Environmental Management and Planning. A range of modelling models from a variety of scientific

326
Different approaches such as GIS, optimisation and AI, simulation disciplines of interest with examples of
categories of modelling, remote sensing, citizen science, and geodesign. their use for NBS)
NBS
Nijhuis et al. Geodesign as a GIS-based planning and design method, It allows project conceptualization,
(2016) which tightly couples the creation of design proposals with impact analysis, design specification,
simulations informed by geographic contexts. It comprises a set of stakeholder participation and
geo-information technology driven methods and techniques for collaboration.
planning built and natural environments in an integrated process

Steinitz Geodesign is proposed as an iterative design method that uses It was shown how geodesign bridge
(2016) stakeholder input, geospatial modelling, impact simulations, and geo-information technology, spatial
real-time feedback to facilitate holistic designs and smart design and planning. It showcases the
decisions. ongoing effort to employ the potential
power of using GIS to link different
model types and ways of designing to
make better plans.

Ecosystem Nelson and Modelling multiple ecosystem services, biodiversity It allows to predict changes in
services Daily conservation, commodity production, and tradeoffs at landscape ecosystem services, biodiversity
provided by (2010), scales using spatially explicit modelling tool, Integrated conservation, and commodity
NBS Nelson et al. Valuation of Ecosystem Services and Tradeoffs (InVEST) production levels. InVEST was applied
(2009) to stakeholder-defined scenarios of
InVEST: land-use/land-cover change in order to
https://naturalcapitalproject.stanford.edu/software/invest help making natural resource decisions
more effective, efficient, and
defensible.

327
7.6 Mimicking the impacts of NBS: how laboratory data can help

Laboratory experiments can help assessing causal relationships based on the


observation of the direct effects of NBS on a small-scale with rapid, and short-
term ecological/environmental processes and society. It can be assumed that if
a laboratory study is well-controlled, the factors that can cause the difference can
be reliably identified. In contrast, all confounding factors cannot be ruled out in
observational studies (Yuan et al., 2017). Thus, laboratory studies are generally
assumed to mimic long-term impacts of NBS and can be useful when trying to
assess ex-ante the performances of a NBS intervention.

For example, a series of laboratory flume experiments has been conducted within
the EU-H2020 Project OPERANDUM in order to study how different soil surface
conditions (smooth, compacted and non-vegetated surface, soil vegetated with
standard herbaceous plants vs specifically selected deep-rooted herbaceous
plants, etc.) may affect or improve the erodibility resistance of the riverbank of
Panaro River (IT) over long term. The studies were antecedent to the actual NBS
deployment and guided the choice of the NBS most appropriate to help preventing
levee failures and inundations at this site.

In the context of research with human, a novel technique to measure the


individual's psychophysiological response to environmental stimuli is represented
by Immersive Virtual Reality (IVR). IVR involves the use of virtual devices that
allow the individual to experience a simulated natural environment in a
multisensory way. For example, the response on the induced stress of virtual
environments at different degrees
of biodiversity (Schebella et al.,
2020) and the aesthetic value and
perception of beauty of a virtual
environment with multiple natural
features (Vercelloni et al., 2018)
can be assessed with this
technique. During the NBS planning
stage, IVR pilot studies could
provide guidance on how to
maximize NBS beneficial effects on
human health and well-being.

7.7 Engaging the community in the data collection process: citizen


science and its role in NBS monitoring

Citizen science has great potential in monitoring and evaluating NBS impact. It
can represent a cost-effective way to gather data on a larger numeric and/or
geographical scale than would otherwise be feasible. In addition to this, citizen
science approaches can offer numerous benefits for society compared to other
types of data generation, including:

328
• Great paybacks to both society and growing areas of science (such as
nature-based solutions), including raising awareness of local risks and
opportunities;

• Engagement and empowerment of the public by giving them a voice in


science, policy and decision making;

• Societal benefits such as social cohesion, integration, and reconnection of


communities with nature.

Citizen science has risen in popularity due to these numerous co-benefits for
citizens. Citizen science-based data generation can also represent added value
for local authorities: Although there can be a cost associated with running such
activities, this can represent value for money compared to the economic cost of
alternative monitoring methods, particularly if the added social benefits are
factored into the ‘value’ of the approaches.

Whilst citizen science approaches


are becoming increasingly
adopted, they can also come with
challenges. This includes
challenges in relation to the quality
of data generated. For example,
evaluation methods may need to
be basic for some indicators
compared to the complexity that
can be achieved through the use of
specialists. Other challenges to
wider adoption of citizen science
projects include the need for
training participants, and
associated problems in retention following training, challenges in validating data
quality and reliability, and eliminating sampling bias (Pocock et al., 2014;
Lukyanenko et al., 2016).

Despite these challenges, citizen science approaches are increasingly being


adopted, including in the evaluation of nature-based solutions. For example,
citizens have been actively involved in data collection for earth observation,
ground measurements, and survey data. Citizens have contributed by using
technological advancements such as smartphones, low-cost sensors, and social
media to record such diverse parameters as air quality, bird and butterfly counts,
water quality, recreational value of greenspaces, and risk management. Data
collected from such processes may represent an entire dataset or can be used as
added value or for validation purpose for data collected using other methods. The
following tools are being successfully and broadly applied for citizen science data
collection.

Crowdsourcing encompasses obtaining a large amount of data from a crowd of


people (or more often the general public) that shares information, voluntarily.
This is often done through the internet and/or using smartphones. Each single
data supplies is then aggregated to generate a cumulative dataset. Due to the

329
large number of contributors, crowdsourcing requires an easy to use framework,
instructions and communication setup to ensure engagement.

Crowd sensed data describes data which are specifically collected and shared
by a large number of citizens through different types of devices, such as mobile
phones, wearable sensors or vehicles (e.g., sensors mounted on bicycles to
measure air temperature or air quality parameters). Whilst this method of data
generation also requires participant permission, it can be less active than
crowdsourcing, with data often collected passively through smartphones and
sensors rather than active input by participants. This can include environmental
factors such as ambient light, noise, location data, movement data, and air
quality. Similarly to crowdsourcing, this method of participatory sensing can
support the monitoring process over a range of spatial scales form small to large
(Guo et al. 2015). It has several advantages such as low-cost sensing or high
amount of data collected. However, the use of crowed sensed data can be
constrained by issues such as sensor accuracy and participation of citizens.

Volunteered geographic information (VGI) is a type of crowdsourced


information where data have spatial information attached. The crowdsourced
data are usually collected in, or converted to, a mapped form with spatial (and
temporal) dimensions. Leading examples for this are OpenStreetMap (OSM) or
the use of online mapping and social media such as Twitter to communicate
information about natural disaster events (e.g., hurricanes and earthquakes).

These and other citizen science approaches have been tested and implemented
by various NBS projects. This includes the EU-H2020 project OPERANDUM in
which citizen science approaches were integrated into the NBS implementation
and monitoring. Indeed, the community neighbouring the NBS were engaged in
the co-design of the nature-based solutions, and were actively engaged in data
co-creation processes. At one of the OPERANDUM NBS sites (Finland), citizens
measured snow depth with traditional and low-cost measurement instruments
during the winter, while water quality and visibility as well as precipitation were
measured throughout the year. The measurements were then shared in a web
application which is linked to the database of the national weather service (where
the data were compared and combined with remote sensing data). OPERANDUM
also uses OSM data to derive information about critical infrastructure for the risk
modelling. Furthermore, the project offers a web application for NBS
crowdsourcing which engages the citizens to post information (through their
mobile phone or the internet) about NBS projects implemented in the place where
they live or more in general about NBS which they have knowledge of (Figure
7-4).

330
(a)

(b)

Figure 7-5. (a) Snapshot of the crowdsourcing app used in the EU-H2020 project OPERANDUM to engage
the community in sharing information on NBS (source: http://crowd-geokip.kajoservices.com/views/map);
(b) citizens involved in the NBS co-deployment and monitoring at Catterline (UK): on the left, residents
helping to measure the permeability of the soil at their front gardens; on the right, residents fixing geo-grid
on slope to prevent erosion and shallow landslides (source: EU-H2020 project OPERANDUM; photo credit:
Alejandro Gonzalez-Ollauri).

7.8 Data integration

In previous sections, different data collection strategies have been explored for
the purpose of fulfilling the data requirements for NBS monitoring and
assessment. Data collection is however only the first steps in conducting a NBS
assessment, since data gathered from different sources will often have to be
analysed in combination and integrated together in order to provide valuable
insights on the impacts and co-benefits of a NBS intervention in comparison to a
baseline scenario.

In that respect, spatial modelling and spatial analysis may represent an effective
strategy for the monitoring and/or planning of NBS, since it allows to integrate,

331
analyse and visualize different data types. For example, using remote sensing
data under a GIS environment, it is possible to provide geo-referenced
information on the shape, size and distribution of different land-use classes of
the urban environment (Herold et al., 2005). This allows monitoring of urban
growth (area change, structures, land consumption, soil sealing) and land
cover/land-use changes (loss of agricultural area, wetland infringement, loss of
areas important for biodiversity, spatial distribution of inner-urban green and
open spaces and natural areas) as well as mapping of various environmental
parameters (data important for urban climate, access to and distribution of open
space, calculation of sealed surfaces).

High resolution remote sensing data


can be combined with measured
pollutant concentrations in a GIS
environment, to map the removal of
PM10 and ozone by urban trees and
estimate the physical removal of
pollutants by trees at specific
locations. Various types of
observations are usually used in
combination with (and/or as input data
of) modelling tools. Besides, results
from 3D numerical models (e.g., Envi-
met model, https://www.envi-
met.com/) and other modelling
techniques can be also usually
imported in a GIS environment and
combined with RS, EO and ground
observations for planning purposes or for analysing present/future impacts of an
NBS intervention. See Table 7-10 for more examples 61.

61
Another relevant example of data integration through digital mapping (e.g., remote sensing, GIS) is
provided in EKLIPSE (http://www.eklipse-mechanism.eu/home)

332
Table 7-10. Examples of data integrations used in NBS projects.

Project Approach Web link

Naturvation Remote sensing, satellite imagery and digital orthophotos together with https://www.naturvation.eu/
Geographic Information Systems (GIS) used to develop a digital elevation
model and a digital surface model. Input data: qualitative and GIS data.
Output data: quality of life, tree coverage; spending time in city parks,
gardens, and open spaces.

Deterministic model which uses remote sensing of greenness as well as


surface sealing to estimate recreation supply. Input data: Remote sensing
data, NVDI and surface sealing. Output data: Spatially normalized
minimum of green space provision per person suggested by the city
administration (m² per Block; m²/m²)

A model based on remote sensing – MODIS NPP. Input data: allometric


equations, net photosynthesis (PSNnet), average growths in diameter of
specific tree species, trees diameter at breast high. Output data: Net
primary productivity kg C per tree and year

IMPRESSIONS Mapping land use, ecosystem functions, and ecosystem services using http://www.impressions-project.eu/
cutting-edge remote sensing and machine learning techniques

A coordinated effort to integrate and analyse a higher quantity and quality


of CO2 and CH4 data, from in situ and remote sensing observations
encompassing atmosphere, land and oceans.

URBES Remote Sensing of Urban Ecology (EO sensors, modelling algorithms) https://www.biodiversa.org/121

Spatial and remote sensing data analyses

333
URBACT Remote sensing (production of high spatial resolution, including the urban https://urbact.eu
atlas, built-up areas, and air pollution) and so-called big data are used to
compare and benchmark cities.

OPERANDUM Remote sensing data to monitor land surface parameters, Observation https://www.operandum-project.eu
from Copernicus Land, Marine, Atmosfere, Climate Change, Emergency
Services, NBS monitoring sensors installations (e.g., monitoring green
roofs in Dublin), GHSL population distribution, EUROSTAT socio-economic
indicators to compute the risk indicators, Local and EU scale hazard
information at corresponding different return levels scenarios and critical
infrastructure as an input to risk modelling, Local and continental ERA40
data reference climate data and CORDEX climate projections to assess
different NBS scenarios for present and future climate.

URBAN Mapping the removal of PM10 and ozone by urban trees by combining high https://www.urbangreenup.eu/about/about.kl
GreenUP resolution remote sensing data with measured pollutant concentrations to
estimate the physical removal of pollutants by trees.

Mapping and assessing the contribution of urban vegetation to


microclimate regulation, deriving a map of Land Surface Temperature
based on Landsat 8 Data, using a model of Du et al. (2015), aggregating
Land types to assess the changes in average temperature.

Mapping urban temperature using remote sensing (split window


algorithm) and modelling techniques for assessing urban temperature and
the indicator for microclimate regulation.

PLUREL Remote sensing and GIS for sustainable urban development science to www.plurel.net
provide geo-referenced information on the shape, size and distribution of
different land-use classes of the urban environment. Main applications:
• Monitoring urban growth (area change, structures, land
consumption, soil sealing;
• Monitoring land cover/land-use changes (loss of agricultural area,
wetland infringement, loss of areas important for biodiversity,

334
spatial distribution of inner-urban green and open spaces and
natural areas);
• Mapping of environmental parameters (base data important for
urban climate, access to and distribution of open space, calculation
of sealed surfaces).

335
Another relevant example of data integration is represented by the use of Big
Data in the context of NBS, where they can be helpful in decoding the complex
relationship of socio-environmental cultural domain. Although there are not yet
well-defined and generalized indices to be used (hence caution should be used in
handling Big Data for NBS monitoring), appropriate measures could be
constructed by combining different data types and data sources, such as (i)
spatial data combined with health data on illness incidence, and (ii) spatial data
on population density and social demographic indicators with a view to analyse
climate change (Frantzeskaki et al., 2019). In that respect, a valuable source of
Big Data is represented by the social media data, which can help identifying new
habits and needs as drivers of uncommon way of life (Ilieva and McPhearson,
2018). Another source of big data is the data generated by consumer behaviour
inspired by sustainable choices. Under this perspective, spatial, economic,
preference and temporal data can be aggregated and analysed.

As further discussed in Section 7.9, the establishment of a baseline also required


the integration of different data types. In this case, spatial data using remote
sensing, earth observation and GIS technologies are usually combined with non-
spatial data from field surveys and other sources if they are secondary data. In
the EU-H2020 project UNaLab, for example, non-spatial datasets including both
qualitative (surveys, questionnaires and scoring, etc.) and quantitative
(environmental, social and economic statistical and legacy datasets) data were
completed with spatial information for the evaluation of KPIs and the
establishment of the baseline conditions.

The non-spatial or attribute or characteristic data typically include demographic


variables, socioeconomic conditions and other non-spatial properties such as
environmental culture or human/individual behaviour (cf. Sections 7.3–7.4).
They are relevant not only for describing the status quo and planning the future
strategy, but for identifying needs too. In the EU-H2020 project URBiNAT, the
well-being, social cohesion and economic-social aspects of the project city have
been analysed through collection of several types of non-spatial data. Other
examples of how various types of non-spatial data can be combined for the
purpose of NBS assessment are provided in Table 7-11.

In some cases, integrated datasets of relevance for NBS monitoring and baseline
construction are also readily available from external sources. An excellent
example is the Global Human Settlement Layer (GHSL) platform
(https://ghsl.jrc.ec.europa.eu/download.php). GHSL produces global spatial
information about the human presence on the planet and its changes over time.
This is in the form of built up maps, population density maps, settlement
classification maps and database on urban centres (see Table 7-12). The
framework uses heterogeneous data including global archives of satellite
imagery, census data, and volunteered geographic information and produces free
information layers and knowledge reporting about the presence of population and
built-up infrastructures at European and Global scales (Pesaresi, 2018).

336
Table 7-11. The use of non-spatial data applied in the NBS projects.

Project Mode of acquisition Main application Source

CONNECTING Gathering knowledge from different stakeholders To identify new synergistic data-gathering https://connecting
Nature through surveys, questionnaires, workshops, reflexing techniques that make use of the latest nature.eu/our-
monitoring webinars and round tables; co-creation available technologies and allow resources
and co-design events with policy makers and the representation of traditionally under-
communities-of-interest; statistical data and policy represented groups in urban policymaking
documents; set of non-spatial human-wellbeing and
economic indicators (e.g., social cohesion, general
wellbeing and happiness, levels of aggressiveness and
violence, additional funding secured for NBS, etc.)

UNaLab Qualitative data (e.g., surveys, questionnaires and To establish the baseline conditions, for https://unalab.eu/
scoring) and quantitative data (environmental, social evaluating the KPIs and complementing the en/documents/d31
and economic statistical and legacy datasets) spatial information with non-spatial attributes -nbs-performance-
and-impact-
monitoring-report

EKLIPSE “Air Quality” indicators developed within the EKLIPSE To assess ecological, economic and social http://www.eklipse
Working Group impact evaluation framework. value of NBS -
mechanism.eu/app
• non-spatial indicators of gross quantities: s/Eklipse_data/we
annual amount of pollutants captured by bsite/EKLIPSE_Rep
vegetation; ort1-
• non-spatial indicators of net quantities: net NBS_FINAL_Compl
air quality improvement (pollutants ete-
produced—pollutants captured + GHG 08022017_LowRes
emissions from maintenance activities); _4Web.pdf

• non-spatial indicators of shares: share of


emissions (air pollutants)
captured/sequestered by vegetation;

337
• the economic value of air or water purification
measured using avoided costs for health care
or replacement costs for artificial treatment

OPERANDUM Surveys on perception of NBS in local communities Asses the acceptance of the NBS by local http://operandum-
communities to provide qualitative input into project.eu
Surveys on implementation of the NBS in the Open-Air efficacy and co-benefits and societal impacts
Laboratories of the NBS. Monitor progress of the NBS
installation to synthetize practical cook-books
of NBS implementation

NATURVATION Urban Nature Atlas (UNA), a database and detailed To assess economic and social value of NBS https://naturvation
characterization of 1000 NBS in 100 European cities; .eu/atlas
set of social indicators identified for the assessment of
NBSs social impacts especially related to well-being
and human health, education, social interaction, social
justice, safety, job creation, urban green space
accessibility and availability

GREEN SURGE on-spatial quality data gathered through interviews, To support decision-making on urban green https://greensurge
questionnaires, and then used in public participation space-management, e.g. to assess how .eu/
geographic information systems (PPGIS) and hedonic residents with different backgrounds value
pricing and use green areas across the cities

Nature4Cities Survey among local residents on how green space can To develop a complimentary assessment tool https://www.natur
contribute to quality of life and also to regional on quality of life, to evaluate the e4cities.eu/results
attractiveness environmental, social and economic benefits
associated to NBS

URBiNAT Survey through validated questionnaires in multiple To assess the level of well-being across the https://urbinat.eu
cities project cities

338
The GHSL database (in particular the Urban Centres Database UCDB) can be used
as data source for assessing several indicators related to SDGs and in particular
the indicators of success of nature-based solutions in cities both at the European
and Global scale. In the EU-H2020 project OPERANDUM, for example, GHSL data
are in combination with hazard information (e.g., flood extent) to derive the flood
risk indicators such as population affected. Another example is the possibility to
use GHSL datasets to investigate changes in the amount of greenness within
cities in the periods centred on the years 1990, 2000 and 2015 (Corbane, 2018).
Of relevance to indicators framework for NBS, GHSL multitemporal dataset on
built-up (GHS-BUILT) and population (GHS-POP) can also be used to provide a
quantitative assessment of changes in the Land Use Efficiency (LUE) indicator for
more than 10 000 cities between 1990 and 2015 (Schiavina et al., 2019). This
measures the land consumption rate to population growth rate and can be used
as a proxy for land take. The LUE is recommended for estimating SDG indicator
11.3.1 which requires data on the spatial extent of the settlements and the
dynamics of their population.

Table 7-12. Summary of main GHSL datasets at global and European Scales. GHSL datasets are described
in detail in Florczyk et al. (2019). All datasets are freely accessible for download from the GHSL website
managed by the European Commission: https://ghsl.jrc.ec.europa.eu/download.php

Dataset Semantic Format Resolution Date Main


input
data
source

GHS- Built-up area and their Raster 30 m– 1975–1990– Satellite


BUILT densities at global scale 250 m– 2000–2015 imagery
1 km

GHS- Density of population at Raster 250m– 1975–1990– GHS-


POP global scale 1km 2000–2015 BUILT
Census
data
GHS-
BUILT
GHS-POP

GHS- Classification of Human Raster 1 km 1975–1990– Census


SMOD settlements: urban 2000–2015 data
centres, urban cluster,
rural areas at global
scale

UCDB Description of spatial Shapefile 1 km Different time GHS-


entities corresponding Excel file depths with a BUILT
to accordingly to a set maximum of
of multi-temporal 40 years GHS-POP
thematic attributes at GHS-
global scale SMOD

339
Other
sources

FUA Functional Urban Areas Shapefile 1 km 2015 GHS-POP


corresponding to urban
centres and their UCDB
commuting zones at Global
global scale friction
matrix

Table 7-12 provides a summary of the main datasets available in the GHLS suite,
which includes, among others, the following data products.

1. The European Settlement Map (ESM_2015) which is a new spatial raster


dataset mapping human settlements of 2015 in Europe. It is published in two
layers: (a) Built-up areas at a spatial resolution of 2 meters, (b) Classification
of the built-up areas into residential and non-residential at a spatial resolution
of 10 meters.

2. The GHS-FUA Functional Urban Areas. This dataset delineates the spatial
entities representing the commuting area of the Urban Centres of 2015 [9].
The dataset is provided in GeoPackage format.

3. The Urban Centres Database (UCDB) in which more than 10 000 individual cities
are characterised by a number of variables (several are mulitemporal) describing
the geography (e.g., temperature, elevation), socio-economic characteristics
(e.g., population density, built-up surface), the environment (e.g., greenness,
CO2 emissions), potential exposure to natural hazards (e.g., exposure to floods,
heatwaves) and SDG indicators The UCDB is provided in the form of vector
shapefiles with attributes describing each spatial entity and in the form of an
excel table with detailed description of each attribute. Furthermore, there is a
dedicate webpage (https://ghsl.jrc.ec.europa.eu/ucdb2018Overview.php) which
allows you to explore the different thematic attributes for each city (Figure 7-5).

340
Figure 7-6. Example of the UCDB visualization for the urban centre of Thessaloniki (GRC) showing the
environmental attributes.

7.9 Baseline Assessment

Baseline data collection is essential for any future evaluation of NBS performance.
Baseline data should essentially be able to convey both the “state of play” (initial
situation, from the social, economic, environmental points of view) as well as
temporal and spatial trends of parameters, which will be further monitored and
assessed throughout the project implementation and at its conclusion. The
assessment is related to the performance evaluation of the NBS itself, and it is
not aimed for the comparison between the NBS intervention and other grey or
hybrid solutions dealing with the same issues. Especially for nature-based
solutions, identifying initial trends allows an understanding of how the baseline
conditions may change in the absence of the proposed actions, and thus for the
definition of “business as usual” scenarios. Baseline data may indicate, for

341
example, that a particular peri-urban
habitat may have significantly shrunk
in the last ten years and is continuing
to shrink at an accelerated rate.
Without an understanding of this
trend, conclusions about the results of
any action and its impact on the
habitat would be erroneous. In fact,
comparing the outcome (e.g., in year
2025) with the initial state (2020) –
rather than with the “business as
usual” scenario (for the year 2025) –
would be flawed in this case.

For physicochemical constituents, the baseline conditions should ideally be


established prior to NBS implementation. In cases when the baseline
measurements are not available, a site with similar conditions could be employed
as a “proxy baseline”. The latter approach naturally has its limitations in the
representativeness as the reference site will not have the same exact conditions,
and the results may be biased. Special regionalization methods could be
employed to minimize the representativeness issues (e.g., selection of multiple
sites with available measurements having similar characteristics to the NBS
implementation site, in order to have a more representative sample). Spatial data
can be employed for assessing the baseline conditions when combined with in
situ measurements. However, historical and statistical datasets may have
variable spatial and temporal resolutions, and they may not be consistent within
a single urban area. Data aggregations or modifications may be necessary to
overcome these challenges in applying the available datasets for pre-NBS
baseline establishment.

Figure 7-7. Key steps in the development of a robust data management plan to ensure data quality, data
standards and data accessibility.

342
The lack of baseline data and/or the fact the baseline data collection is not always
envisaged in an NBS project and often depends on (Bamberger, 2006):

• Lack of awareness on the importance of baseline data for NBS impact


assessment;

• An inadequate and insufficient program planning and oversight;

• Budget/Time/ Political constraints;

• Delays in the administrative procedures (recruiting and training of the


staff, acquisition of the necessary materials, commissioning consultants
etc.) before the beginning of the baseline study;

• Evaluation not commissioned until late in the project cycle;

• Difficulties in identifying common data groups for the comparison;

• Lack of availability or low granularity of initial data.

Table 7-13 provides general guidelines on how to determine whether a baseline


study is necessary, and to what extent (International Federation of Red Cross
and Red Crescent Societies, 2013).

Table 7-13. Necessity of baseline data studies (based on the guidelines provided by the Planning and
Evaluation Department (PED) of the International Federation of Red Cross and Red Crescent Societies).

Baseline study Rationale

No study needed Sometimes it is not necessary to study and collect baseline data because
they are already known, e.g.:
• The indicator value may be known to be “0” prior to the project
start (for instance “none of the communities have been
involved in NBS co-implementation before the project”).
• The data could be available from other sources (i.e., from
secondary data).

Shallow Study The number of baseline data and the methods to measure them are
Needed restrained in time, capacity and resources because they are available
from other sources, therefore easily collectable, or it possible to replace
expensive household surveys with less costly qualitative methods such
as individual / group interviews or online surveys. For example,
“Perceived neighbourhood green space safety”, assessed via individual
questionnaires using random sampling techniques.

In-depth Study Sometimes, it is necessary to have a more rigorous baseline study.


Needed Examples could be climate resilience improvement projects in which it is
foreseen a renovation of the buildings' roofing, that could require the
collection of data regarding energy and carbon emissions savings (i.e.,
from reduced building energy consumption (kWh/y and t C/y saved)),

343
the development of specific questionnaires to be submitted to residents
and a statistical analysis of the data.

Reconstruction When the baseline data study is needed but it was not conducted prior
of Baseline Data or near to the project beginning, a reconstruction of the baseline
measurements is needed. The greater the time lag between the delivery
of the project activities and the baseline study, the more likely the
project will have a measurable effect on the indicators, leading to an
underestimation of its impacts on the context.

Nevertheless, assessment of project outcomes and impacts should not be


confined strictly to baseline and final analyses, because NBS projects may yield
cascading results or externalities during the actual implementation. For example,
in the cases in which a project implies an improvement of the green areas present
in the neighbourhood, speculators may begin to invest in land ownership and
families can decide to start improving the quality of their property. If the baseline
data study is postponed for a long time, many of these important changes could
be omitted.

If baseline data need to be reconstructed, there are several approaches which


can be used to achieve a discreet result (Bamberger, 2010):

• Secondary data: checking documentary sources, such as annual reports


of governmental agencies:

• Administrative data: feasibility and planning studies made prior to an


intervention on a specific territory, application / registration forms, etc.;

• Recall: technique based on surveys or individual / group interviews,


particularly useful for recalling major events or impacts of a new service
(including ecosystem service), albeit subject to biases;

• Key informants: in-depth interviewing and involvement of external


stakeholders (representatives of a society or a specific target group)
which combines “factual” information with a particular point of view,
offering a different perspective.

It should be, however, noticed that no data collection method is free from the
possibility of inaccuracy. Due to this, the above-mentioned methods, and
especially the ones relying on surveys and interviews, are usually accompanied
by the Triangulation method. This allows to verify the results against data
collected from other sources, to confirm accuracy and precision of the
reconstructed baseline. Another term often encountered in baseline studies is
Comparison (or Control) Group. It refers to a group of units (e.g., persons;
census cells; households) that has not been affected by the project impacts and
serves as a source of counterfactual causal inference (Maldonado and Greenland,
2002). The big challenge in this case is selecting a well-matched baseline
comparison group.

344
A critical point whose importance is sometimes overlooked is the fact that spatial
analysis of data for baselines requires a priori knowledge about both the data as
well as the underlying processes (Csillag and Boots, 2005). This includes being
aware of the possibilities and limitations of the various spatial statistics available,
but also knowledge of existing urban policies, spatial plans and regulations which
allow contextualization of findings.

Baseline studies, for example for Strategic Environmental Assessment (SEA,


https://ec.europa.eu/environment/eia/sea-support.htm), include reviews of the
policy context and a collection of detailed evidence on the state of the
environment (context) in which a nature-based solutions project will deploy its
activities.

Lack of statistical data can hinder the creation of a sufficiently robust baseline
profile for one or several key NBS assessment domains, potentially leading to a
limited understanding of pre-conditions and potential. One way of mitigating this
risk, tested in proGIreg, was to include a “long list” of spatial indicators, ensuring
that even though cross-city comparability may be limited, the key assessment
topics are still characterised by a minimum of two data sets, selected from the
most commonly used datasets of statistical offices across Europe. These have
been grouped in key assessment domains, and descriptors (Table 7-14), with
each descriptor further expanded through a set of indicators and datasets – 70
in total.

Based on the proGIreg experience, there are two recommendations which can be
provided for the purpose of developing baseline analyses. The first is the
allotment of sufficient time for data collection, as a task in itself which often
involves sending out data requests to other institutions (e.g., regional offices).
Beyond data availability, a key factor of success is the capacity of the cities
themselves to work with data, and the need for close connection between
different stakeholders involved in data management, analysis, policy makers, and
the local communities (as both beneficiaries as well as data providers). This is
likewise a process which should be planned carefully in time.

Table 7-14. Example of baseline data requirements (from EU-H2020 project proGIreg. More details can be
found in Leopa and Elisei, 2020).

Assessment Subdomain/descriptor and example data


domain

1. Socio-Cultural 1.1 Demographics (e.g., Population growth rate, migration rate)


Inclusiveness

1.2 Social and cultural inclusiveness (e.g., Material deprivation rate)

1.3 Education and access to social and cultural services and


amenities

1.4 Housing (e.g., Density of the built environment)

345
2. Human health 2.1 Health (e.g., Incidence of cardio and respiratory diseases,
and wellbeing obesity rate)

2.2 Wellbeing (e.g., Green space per capita, urban safety)

3. Ecological and 3.1 Land use and Vegetation


environmental
restoration
3.2 Climate/Meteorological data

3.3 Air Quality

3.4 Soil

3.5 Water

3.6 Urban environment

4. Economic and 4.1 Market labour and economy indicators (e.g., Number of green
labour market jobs)
benefits
4.2 Gentrification indicators (e.g., Average household disposable
income, property values)

4.3 Tourism and attractiveness indicators (e.g., Expenses in local


retail business)

4.4 Taxes, Investment and Financing (public investment programs)

7.10 Data adequacy and related aspects

Adequate collection, management and use of data is foundational for a holistic


assessment of NBS performances. Challenges and requirements related to data
needs and their collection addressed in the previous Sections emphasise the
importance of generating reliable data. Table 7-15 lists the principal aspects
determining the quality of analysis derived from the main data collection and
generation methods in terms of potential error sources and their prevention and
elimination.

This section focuses on the most


common and critical challenges
encountered when using data. Data
utilization challenges generally fall
into three categories: data quality,
data appropriateness and data
accessibility. Gaps and irregularities
in spatial and/or temporal data
series, as well as data accuracy and

346
other error sources, affect the quality of data, while data granularity and resolution
define if a dataset is appropriate with respect to the target of investigation. Together
with accessibility and other key characteristics discussed at this end of this section,
these aspects determine the overall adequacy of a dataset.

347
Table 7-15. Data accuracy, typical errors and ways to prevent errors for different NBS data generation and collection methods.

NBS data Data accuracy Typical errors/biases Ways to prevent errors


collection/
generation
method

Observational Depends greatly on data Manual sampling data can contain Standardized sampling methods and protocols,
data collection or generation methods, uncertainties due to spatial and appropriate measuring intervals, detection
(Sections e.g., granularity and resolution of temporal heterogeneities or low- limits and calibration of the measurement
7.1-7.2) the measurements, quality of quality measurements. instruments (e.g., Pepper, Brusseau, and
measurement systems, Artiola, 2004).
measurement scale or Random selection of samples may
specification, and selection of cause inaccuracies. Accurate baseline or reference definition.
samples.
Inadequate baseline or reference Statistical manipulations, such as aggregation
definition. (scaling-up) of dis-aggregation (downscaling)
of datasets with varying granularity, must be
Ambiguous or erroneous results when exercised cautiously (e.g., Scholes et al.,
aggregating historical or legacy 2013)
datasets with observational data (e.g.,
Scholes et al., 2013). Satellite observations must be validated
against and complemented by ground
Satellite-derived images can contain measurements and/or other high-resolution
shadows due to the size of the frame RS-platforms such as drones or aircraft-based
or be of low spatial and temporal (e.g., Orgiazzi et al., 2017).
resolution.

Surveys and Survey data are usually collected Poor representativeness or small size Choosing data collection sources/methods
census from a group of participants of a research group. Data from which produce the desired information.
(Sections 7.1 which will represent a larger qualitative survey can be complex.
and 7.3) group. Accuracy of the data In quantitative surveys, verifying quality,
depends e.g. on the Constrains and limitations in relevance, simplicity, accuracy and clarity of
representativeness of the availability of specific or updated the questionnaire.
participant group and sample statistical data.
size. Statistical analysis can be In qualitative surveys, choosing proper
used to estimate the accuracy. approach and identifying suitable strategies for
data collection.

348
Laboratory Laboratory experiments can Samples are not representative for the Verification of the methods to mimic real-life
experiments control most of the variables desired research subject (e.g., situation and long-term effects.
(Sections under study and can offer the samples are not in their natural state)
7.1and 7.6) most accurate analysis methods. or the laboratory experiment is not Well-controlled, standardised measurements
Representativeness of the mimicking the real-life situation or with high-quality and calibrated instruments.
samples and quality of the long-term effects.
analysis define accuracy of the Automated analysis can eliminate human
data. Instrumental or human errors. errors.

Numerical Models are simplifications of the Limitations and uncertainties related Use of high-quality models to address the
simulations real-world systems (Grützner, to the technical or mathematical specific, desired research questions
and 1996) and some uncertainty structure on which the model is built.
modelling should be accepted. Models are calibrated and verified against
(Sections 7.1 Inadequate calibration and/or observational (field or laboratory) data to
and 7.5) The accuracy of the model validation due to low-quality or limited ensure the accuracy of results and the overall
depends mostly on the amount of initial data. uncertainty.
accuracy and of the initial data,
quality of the model, and skills of Inaccurate assumptions and/or Sensitivity analysis performed for the
the model user (Government of approximations in the model. parameters in the model (Government of
South Australia, 2010). South Australia 2010).
Inappropriate use of the model.

Citizen In complex data collection Sensor accuracy is too low. Clear protocols, frameworks, and instruments
science methods, variability of data including those for transparent communication
(Sections 7.1 collected by volunteers as non- Too complex data collection methods (e.g., Dickinson and Bonney, 2012; Dickinson,
and 7.7) professionals can be greater for unexperienced users. Zuckerberg, and Bonter, 2010).
compared to professionally
collected data (Aceves-Bueno et Instructions are misunderstood. Proper training of the volunteers (e.g.,
al., 2017). However, citizen Dickinson and Bonney, 2012).
science can offer broader Challenges with validating the data
collection of data, analysis of the quality (Pocock et al., 2014) Adopt more advanced statistical analyses to
data accuracy is required in identify errors (Dickinson, Zuckerberg, and
citizen science projects. Bonter, 2010; Pocock et al., 2014).

Collection of a greater number of samples to


eliminate sampling bias (e.g., Gardiner et al.,
2012)

349
7.10.1 Data gaps and irregularities

In many cases, data gaps exist in monitoring efforts. Data gaps can be spatial or
temporal. Also, low quality of the data can be considered as insufficient data
collection. Data gaps can exist in all types of monitoring, including manual or
automated measurements, surveys and questionnaires. Often, when the
monitoring plan is built, the main aspects to be considered are the frequency of
monitoring and distribution as well as the amount of monitoring sites. This is
because data gaps are mainly caused by data provision interruption or insufficient
observation coverage (both sampling frequency and spatial distribution). This
data gaps may lead to data insufficiency which can disqualify the dataset from
the holistic NBS performance assessment. Insufficient data collection may also
originate from the lack of resources. In the interpretation of the monitoring
results, it is critical to identify the data gaps. There are existing techniques to fill
the data gaps e. g. spatial/temporal interpolation, but a special attention should
be paid in order not to degrade the representatives of the data. Table 7-16 lists
the data gaps identified by some of EU-H2020 projects on NBS.

Table 7-16. Data gaps identified in EU-H2020 projects (selected examples).

Project Identified data gaps

ConnectingNature Indicator data are foreseen to cover less than 50% of the Connecting
Nature core indicator list. Therefore, there is a requirement for further
rounds of identification of suitable data sources to be undertaken, and
there may also be a need for new observations and site surveys to be
undertaken to fill in any gaps.

proGIreg Gaps in statistical data due to:


• No cities have been able to provide all the requested data
• Depending on the city, some data are not available on a yearly
base

OPERANDUM • Lack of hydro-meteorological observations time series/low


station density which was partially resolved using remodelled
ERA40 data set
• Gaps in in-situ meteorological observations

Inala • Some cities are not able to expose NBS monitoring data
• Baseline data for some of the NBS are missing
• During the monitoring period, there is a risk of gaps and time-
series inhomogeneity (e.g., precipitation, air quality)

Urbina The project involves and compares several European cities in order to
develop sustainable health corridors. However, the availability of
socioeconomic official data differs from city to city

As an example of how to analyse existing data gaps in monitoring, the California


Department of Water Resources (2016) presents a data gap analysis flow chart
for groundwater monitoring. First question when planning a data collection

350
procedure is to consider the needed types of data, for instance water level and
water quality. It should be then considered if the quantity and quality of the data
are sufficient. After this, data gaps are identified. As mentioned, the data gaps
can be spatial, temporal, or they can be related to low data quality. Temporal
data gaps are related to insufficient frequency of the monitoring, and spatial data
gaps to insufficient number of monitoring sites. As an example of low quality
data, it can originate from insufficient collection or data management methods.
After identification of the data gaps, causes for the existence of the gaps should
be identified. The causes can be related to insufficient funding and resources but
also to insufficient access to the data. Actions to reduce the amount of data gaps
are to increase density, frequency, and quality of the monitoring.

7.10.2 Data granularity and resolution

Data granularity is one of the most critical parameters for successful evaluation
of NBS performance and impacts, because it allows to define an effective and
efficient solution, or (if not well dimensioned) can impede the achievement of the
goals of a project. Data granularity indicates the level of detail expressed by each
single part in a dataset. Different granularities indicate different levels of
aggregation in the dataset. Examples of aggregation include:

• Temporal aggregations: year, month, minute, second

• Distance aggregations: kilometres (km), hectometres (hm), decametres


(dam), metres (m), decimetres (dm), centimetres (cm), millimetres
(mm)

• Geographic (or zonal) aggregations: world, continent, country, city,


district, street, address

• Video aggregations: HD, FULL HD, 4K, 8K

Fine-granularity (low level of aggregation) provides more details than coarse-


granularity (high level of aggregation) making it more helpful for decision-
making. In fact, the higher amount of information ensured by fine-granularity
permits to better target the problem to be solved (i.e., climate change, social
issue, service inefficiency), by making the correlation between causes and effects
more comprehensive.

Since a variety of data types (collected with a likewise variety of monitoring


methods) are required to obtain a full NBS assessment in all its dimensions
(ecological, social, etc.), it is imperative that the granularity of all the different
datasets matches the scale of main driving processes behind the NBS and the
impact of NBS interventions. A reliable evaluation cannot otherwise be obtained.
As an example, in the EU-H2020 project proGIreg problems in data granularity
were encountered due to the different scales at which statistical data are available
in the different countries, and due to the small size of most of the implemented
NBS with respect to the scale available for statistical data.

351
Unfortunately, a general formula for defining the granularity level does not exist.
Thus, technical designers can leverage only on their good experience to set the
correct aggregation in the range of fine-grain and large-grain, considering the
variability of the monitored phenomenon and the level of detail needed for the
evaluation and eventually the use of proxy variables to improve the granularity
of the main variable. Table 7-17 shows the possible levels of data granularity
required to evaluate the impact of an NBS for some specific examples.

Table 7-17. Examples of adequate vs inadequate data granularity levels.

Topic Goal of the Study Adequate/Possible Inadequate/Wrong Data


Data Granularity Granularity

Urban Heat Assess daily • Fine grain: 30 • Over sampled:


fluctuations of the minutes second,
urban temperature • Medium grain: 60 millisecond
minutes • Lower sampled: at
• Coarse grain: 180 day scale no
minutes changes can be
observed

Flooding Assess flooding • Fine grain: day • Over sampled:


events per year • Medium grain: 5 minute, second,
days millisecond
• Coarse grain: 30 • Lower sampled: at
days year scale no
changes can be
observed

Urban Green Estimate green • Fine grain: 10 • Over sampled:


Areas density in the sq.m (*) sq.cm, sq. mm (*)
urban area • Medium grain: • Lower sampled:
200 sq.m 30 sq.Km
• Coarse grain: 1
sq. Km

Urban Assess yearly • Fine grain: • Over sampled:


Transportation fluctuations of number of number of
users of urban passengers at 30 passengers per
transportation minutes (for each second (for each
line) line)
• Medium grain: • Lower sampled:
number of number of
passengers per passengers per
day (for each line) year (total)
• Coarse grain:
number of
passengers per
month (total)

(*) sq. stands for square. For example, sq.m stands for square metre.

352
When talking about a representation (e.g., video streaming, image, photo, spatial
data), granularity takes the name of resolution and indicates the size of the
minimum unit/area in a representation (e.g., video streaming, image, photo,
spatial data). Spatial resolution is a common and essential feature in monitoring
systems and indicates the ability of the sensor to detect details of the complex
environments, and the minimum area is measured in meters.

Low spatial resolution sensors (30–300 m) produce adequate results at large


scales, although they are incapable of capturing greater amount of details as high
spatial resolution outputs (less than 30 m). High resolution is essential for
characterisation and interpretation of complex environments and models. As
example, in urban flood and hydraulic studies of river and floodplain interactions,
topographic details significantly influence the capability to discover the flow path
interactions with the underlying terrain (Krebs et al., 2014; Mason et al., 2007).

In conclusion, to assess the impact of a NBS or the development and distribution


of a phenomenon in the ecosystem, it is critical to define the correct level of
aggregation, the data granularity, of the measurements for both the time
(temporal granularity) and the location (spatial resolution). In that respect, fine-
grain and high-resolution local monitoring sensors (or their combination) often
represent the most suitable option to record the actual changes in the urban
system fostered by the implemented NBS.

7.10.3 Data Accuracy

The accuracy is the qualitative parameter indicating the degree of correctness


of a measure derived from the direct observation (sample) with respect to the
objective true or the reference value. In other words, the accuracy quantifies how
much a measure is near the actual value. The common way to express the
accuracy is the percentage, calculated with respect to the full scale of the sensor,
or with respect to the sample. As example, a temperature sensor with full scale
of +-50° and accuracy of +-1% (+-0.05°), means that with an actual value of
30° the sensor could produce a measure in the range between 29.95° and 30.05°
(Figure 7-6).

Figure 7-8. Temperature sensor with full scale of +-50° and accuracy of +-1% (+-0.05°): The sensor can
produce a measure in the range between 29.95° and 30.05, if the actual value to be measured is 30°.

353
Another relevant qualitative parameter in the context of monitoring activities is
the precision that indicates the degree of convergence (or dispersion) in a
collection of samples. In other words, precision indicates how much independent
samples are near among them. The precision is strictly dependent from the
effectiveness of the combination of sensors adopted and methodologies
implemented during the observations. In fact, despite each sensor expresses
static qualitative performance, the combination of sensors with different
methodologies could produce different precision and vice versa.

To better clarify the relationship between precision and accuracy, Figure 7-7(a)
represents the results obtained with a good quality temperature sensor. That
sensor has high precision and high accuracy and for each observation collects
measures aggregated near the actual value. Figure 7-7(b) represents the results
obtained with a temperature sensor with high precision and low accuracy that for
each observation collects aggregated measures, but far from the actual value.
Figure 7-7(c) represents the results obtained with a low quality temperature
sensor with low precision and low accuracy that for each observation collects
measures dispersed and far from the actual value.

Figure 7-9. Measurements obtained with a temperature sensor which has (a) high precision and high
accuracy (good quality sensor); (b) high precision bur low accuracy; (c) low precision and low accuracy (low
quality sensor). The red dot represents the actual (“true”) value of temperate. The blue and green dots
represent the first sample collection (Observation 1) and second sample collection (Observation 2)
respectively.

Accuracy and precision are critical qualitative parameters to be taken into account
during the monitoring activities. In fact, they indicate the quality of data and, as
a consequence, are decisive to approve or reject the models and related

354
elaborations that are the base line for supporting the performance monitoring,
impact assessment and more in general the decision making.

7.10.4 Biases, main error sources, and data reliability

Aggregation and resolution provide useful information about the dimension of the
measures. However, the observations can be influenced by uncontrollable and
predictable factors that can introduce accidental and systematic errors that could
invalidate the sampled measurements.

Accidental errors are caused by unpredictable conditions (as lack of energy or


connection, vibrations near an instrument, wind) that randomly influence the
results and for this reason they cannot be avoided.

A bias is a systematic error that introduces a constant or proportional deviation


(absolute or percentage) with respect to the actual value. Biases can be
generated by different unfavourable conditions:

• Instrumental: inadequate, out of scale, or not well calibrated sensors;

• Methodology: approximated models, incorrect formulas and elaborations,


inadequate experimental conditions (e.g., temperature, humidity, not
appropriate insolation);

• Personal: lack of expertise in the operator, parallaxes, interferences,


improper use of the sensors or the methodology.

Despite the accidental and systemic errors cannot be eliminated, a good and
complete monitoring plan will permit to prevent and identify potential conditions
that could generate errors. Identified errors can be solved or minimised with the
application of the corrective actions, such as identification of the incorrect
samples, definition of more precise methodologies, procedures and rules.

Error sources:

• Not identified and corrected systematic error;

• Lack of attention, or overload of work;

• Overlaps applying heterogeneous methodologies or procedures.

7.10.5 Data Accessibility

Quantitative and qualitative data generated throughout the NBS monitoring


periods via remote and in-situ observations, questionnaires, surveys or other
means may have different access rights (e.g., open, semi-open, or confidential)

355
depending on the degree of confidentiality originally specified in the legal or data
management plans. It may be openly available or subjected to access restrictions
imposed by governing bodies or EU-level regulations, such as General Data
Protection Regulation (GDPR) (EC, 2016). The latter concerns the personal data
collection during, for example, Urban Living Lab (ULL) sessions, health and well-
being surveys or other studies involving humans. Naturally, not all data
generated can be made public, so any personally identifiable information, which
can be potentially generated during the project, should be carefully considered
before and throughout NBS implementation to avoid disclosing any sensitive
information. Here, it should be noted that availability and accessibility mean
“existence” and “possibility and ease of retrieval”, respectively. While accessible
data is concomitantly available, “availability” does not imply “accessibility”.

Although municipalities or other data owners may be reluctant to make their data
open access and share this data with the third parties, open data has numerous
benefits over restricted access data. Often, numerous datasets do not bring any
additional value because of their inaccessibility to the third parties. Open data
can be widely utilised by research institutes and universities by applying it in
research and education to generate, for instance, projections and scenarios based
on the historical records. The possibility to use open datasets for producing
various simulations and utilising
them for NBS baseline conditions
assessment brings an added value
to the datasets and their owners.

Data accessibility plays a critical


role in establishing a holistic NBS
evaluation framework as it is
essential for establishing pre-NBS
baseline conditions. When only
fragmented or irregular datasets
are accessible, it creates
considerable bias and the possible
need for data aggregation or other
modifications of data points leading to biased outputs. In that respect, caution
should be exercised when, for example, EU-wide datasets available from external
sources are integrated in the NBS monitoring framework.

Despite the restricted access to some of the datasets being generated during the
NBS projects, many data and results are accessible through the platforms
established by the projects. This is of outmost importance as data-informed
decision- and policymaking are critical for a wider NBS implementation in urban
areas. Not only open data provides such attributes to urban development, it
encourages greater collaboration in NBS implementation through ample evidence
of benefits and issues recorded and obtained via open data sources.

356
7.10.6 Metadata and data standardization

The increasing effort in providing science-based evidence of NBS effectiveness


and (co-)benefits has resulted into increasing volumes of data required for and/or
connected to the monitoring and assessment of NBS interventions. These data
are often associated with single-case studies and disseminated to a small
community (usually the group of main investigators involved in a given NBS
project), but no established protocols are yet in place that guarantee their
accessibility and long-term re-usability by the large community. This clearly
undermines our ability to achieve statistically meaningful evidence and more
generalizable results on NBS performances and impacts, besides impeding the
possibility to take full advantage of data which already exist but are either not
accessible or easy to understand.

It therefore of crucial importance that NBS-related data become aligned to FAIR


data principles, following the example of other disciplines and research fields.
FAIR is an acronym which stands for Findability, Accessibility, Interoperability,
and Reusability of data (Wilkinson et al., 2016). These four principles have been
endorsed by the EC (Hodson et al., 2018) and many other institutions worldwide
as those that should guide the design and implementation of any good data
management, in order to ensure and maximize digital data discoverability and
exchange. In practice, this means that NBS data producers and publishers should
make an effort in following the guidelines (FAIR principles) summarized in Figure
7-8a or, in simpler words, that NBS data should be supplemented by contextual
documentation, provided with persistent identifiers and metadata, and common
standards adopted for both data and metadata (Figure 7-8b). In this perspective,
metadata are an essential aspect of data standardization.

Metadata, or data about data, enrich dataset with additional information such
as basic characteristics of the datasets (e.g., measured phenomena, author, and
spatial/temporal resolution), quality, and completeness. This allows users or
computers to better assess datasets for a specific use. Metadata enable easier
data discovery since it exposes information about the data which would normally
be hidden within the dataset itself. This allows inspecting information such as
quality, resolution or spatial/temporal coverage without opening/inspecting the
dataset and allows seamless integration of data from different sources. Several
International standards exist which facilitate an easier adoption of FAIR
principles.

357
Figure 7-10. (Top) The set of Fair Principles (source: Wilkinson et al., 2016). (Bottom) a simplified schema
explaining the key elements needed to ensure FAIR data (source: Hodson et al., 2018).

358
Table 7-18 lists some of the most relevant standards in the domain of geospatial
data, metadata and services. Another example of standard is the EU Directive
INSPIRE (https://inspire.ec.europa.eu/), which aims to create a European Union
spatial data infrastructure where environmental data collected on a national basis
can be shared and used on a pan-European basis. In recent years, the importance
of data standardization has become clear also in the context of NBS and some
NBS projects have made significant efforts in developing successful data
management plans. For example, the EU-H2020 project OPERANDUM has
developed a NBS data portal which is fully compatible with semantic web and is
OGC and INSPIRE complaint. Data newly generated by the project (along with
data gathered from external sources semi-automatically) are complemented with
metadata and harmonized according to ISO standards, thus fulfilling the FAIR
principles. For more information on FAIR recommendations and guidelines, the
reader can refer to the EC report by Hodson et al. (2018). For examples of
international standards applied in the context of NBS projects, see Vranic et al.
(2019).

7.11 Conclusion

Successful evaluation of NBS performance and impact rely on the selection of the
appropriate data collection methods, and the quality of data and its inherent
characteristics (e.g., granularity and homogeneity) generated throughout the
NBS monitoring period. This Chapter covered a variety of data types and data
acquisition and generation techniques and discussed their benefits and limitations
applicable to NBS impact evaluation.

Information for NBS impact evaluation, including a crucial step of baseline


assessment, can be obtained via multiple sources, including in-situ
measurements, laboratory experiments, remote sensing or Earth observation
techniques, and citizen science. The selection of data collection methods should
be based on solid planning, technical expertise, and a wide knowledge of the
state of the environment and its functioning to ensure that the relevant and
accurate data are collected for the purposeful NBS monitoring and assessment.
Current and projected NBS impact can be further evaluated by modelling. All data
produced during the NBS monitoring activities must undergo careful evaluation
for possible biases and main error sources to ensure its adequacy and reliability.

Data collection and generation methods for NBS impact assessment discussed
herein can be supplemented with a multitude of datasets obtained from the inter-
European and international databases, although special care should be taken
regarding their spatial and temporal resolution. Collected and generated data
from a variety of sources can be integrated to provide valuable insights on the
impacts and co-benefits of a NBS intervention in comparison to a baseline
scenario.

Examples from the NBS projects regarding, for instance, non-spatial and spatial
data integration, data gaps and modelling approaches to complement data
generation were highlighted throughout the Chapter.

359
Table 7-18. Relevant International data standards following ISO, OGC, etc.

Category International Standards Description

Observations ISO 19156 A conceptual schema for observations, and for features involved in sampling when making
(Observations and observations. It provides models for the exchange of information describing observation acts
Measurements) and their results, both within and between different scientific and technical communities.

SensorML (OGC Sensor It provides a robust and semantically-tied means of defining processes and processing
Model Language) components associated with the measurement and post-measurement transformation of
observations.

SOS (OGC Sensor It defines a web service interface which allows querying observations, sensor metadata, as
Observation Service) well as representations of observed features. Also, this standard defines means to register new
sensors and to remove existing ones

SPS (OGC Sensor It defines interfaces for queries that provide information about the capabilities of a sensor and
Planning Service) how to task the sensor.

STA (OGC SensorThings It provides an open, geospatial-enabled and unified way to interconnect the Internet of Things
API) (IoT) devices, data, and applications over the Web.

Geospatial Data ISO 19107 (Spatial Conceptual schemas for describing the spatial characteristics of geographic features, and a set
schema) of spatial operations consistent with these schemas.

ISO 19125 (Simple A simplified model of ISO 19107 which consists of two parts: 1) a common architecture for
feature access) geographic information, and 2) a specific Structured Query Language (SQL) schema that
supports storage, retrieval, query and update of simple geospatial feature collections.

ISO 19136 (Geography An XML encoding in accordance with ISO 19118 for the transport and storage of geographic
Markup Language) information modelled in accordance with the conceptual modelling framework used in the ISO
19100 series of International Standards and including both the spatial and non-spatial
properties of geographic features.

360
ISO 19129 (Imagery, Framework for imagery, gridded and coverage data. This framework defines a content model
gridded and coverage for the content type imagery and for other specific content types that can be represented as
data framework) coverage data.

Metadata ISO 19115 (Metadata) It defines the schema required for describing geographic information and services by means of
metadata. It provides information about the identification, the extent, the quality, the spatial
and temporal aspects, the content, the spatial reference, the portrayal, distribution, and other
properties of digital geographic data and services.

ISO 19139 (Metadata It defines XML based encoding rules for conceptual schemas specifying types that describe
XML schema geographic resources. The encoding rules support the UML profile as used in the UML models
implementation) commonly used in the standards developed by ISO/TC 211. The encoding rules use XML
schema for the output data structure schema

Services ISO 19119 (Services) Platform requirements on how services shall be created, in order to allow for one service to be
specified independently of one or more underlying distributed computing platforms.

ISO 19128 (Web Map Specifications on the behaviour of a service that produces spatially referenced maps
Server) dynamically from geographic information.

ISO 19142 (Web Specifications on the behaviour of a web feature service providing transactions on/access to
Feature Service (WFS)) geographic features in a manner independent of the underlying data store. It specifies
discovery operations, query operations, locking operations, transaction operations and
operations to manage stored parameterized query expressions.

OGC WCS (OGC Web Specifies the behaviour of a service that serves multi-dimensional coverage data. WCS Core
Coverage Service) specifies a core set of requirements that a WCS implementation must fulfil.

OGC CAT (Catalogue Catalogue services support the ability to publish and search collections of descriptive
Service) information (metadata) for data, services, and related information objects. Metadata in
catalogues represent resource characteristics that can be queried and presented for evaluation
and further processing by both humans and software.

361
Common ISO 19103 (Conceptual Rules and guidelines for the use of a conceptual schema language within the context of
Conceptual schema language) geographic information. The chosen conceptual schema language is the Unified Modelling
Model Language (UML).

ISO 19109 (Rules for Rules for creating and documenting application schemas, including principles for the definition
application schema) of features.

ISO 19118 (Encoding) Requirements for defining encoding rules for use for the interchange of data that conform to
the geographic information in the set of International Standards known as the "ISO 19100
series".

362
7.12 References

Aceves-Bueno, E., Adeleye, A.S., Feraud, M., Huang, Y., Tao, M., Yang, Y. and Anderson, S.E., ‘The Accuracy of
Citizen Science Data: A Quantitative Review’, Bulletin of the Ecological Society of America, Vol. 98, No 4,
2017, pp. 278-290. https://doi.org/10.1002/bes2.1336

Arnold, J.G., Kiniry, J.R., Srinivasan, R., Williams, J.R., Haney, E.B., and Neitsch, S.L., Soil and Water
Assessement Tool: Input/Output Documentation, Texas Water Resources Institute, 2012.

Bagheri Z., Nadoushan, M.A., and Abari, M.F., ‘Evaluation the effect of green space on air pollution dispersion
using satellite images and landscape metrics: A case study of Isfahan City’, Fresenius Environmental
Bulletin, Vol. 25, No 12, 2017, pp. 8135-8145.

Baldacchini, C., Monitoring and Assessment Plan, Deliverable No. 4.1, proGIreg. Horizon 2020 Grant Agreement
No 776528, European Commission, 2019, pp. 124.
https://progireg.eu/fileadmin/user_upload/Deliverables/V2_D4.1_proGIreg_CNR_2020-06-
13_amendment_02_.pdf.

Bamberger, P., ‘Reconstructing Baseline Data for Impact Evaluation and Results Measurement’, PREM notes, No
4, The World Bank, 2010. http://hdl.handle.net/10986/11075

Bamberger, P., Rugh, J. and Mabry, L., Realworld Evaluation: Working Under Budget, Time, Data, And Political
Constraints, Sage Publications, Thousand Oaks, 2006.

Baró, F., Chaparro, L., Gómez-Baggethun, E., Langemeyer, J., Nowak, D.J., Terradas, J., ‘Contribution of
ecosystem services to air quality and climate change mitigation policies: The case of urban forests in
Barcelona, Spain’, Ambio, Vol. 43, 2014, pp. 466–479.

Barton, H., and Grant, M., ‘A health map for the local human habitat’, Journal of the Royal Society for the
Promotion of Health, Vol. 126, 2006, pp. 252–253. doi: 10.1177/1466424006070466

Bass, B., Krayenhoff, S., Martilli, A., and Stull, R., Mitigating the urban heat island with green roof infrastructure,
Urban Heat Island Summit, Toronto, 2002.

Bealey, W.J., McDonald, A.G., Nemitz, E., Donovan, R., Dragosits, U., Duffy, T.R., and Fowler, D., 2007.
‘Estimating the reduction of urban PM10 concentrations by trees within an environmental information
system for planners’, Journal of Environmental Management, Vol. 85, 2007, pp. 44–58.

Beyer, K.M., Kaltenbach, A., Szabo, A., Bogar, S., Nieto, F.J. and Malecji, K.M., ‘Exposure to neighborhood green
space and mental health: evidence from the survey of the health of Wisconsin’, International Journal of
Environmental Research and Public Health, Vol. 11, 2014, pp. 3453-3472.

Bojinski, S., Verstraete, M., Peterson, T. C., Richter, C., Simmons, A., and Zemp, M., ‘The concept of essential
climate variables in support of climate research, applications, and policy’, Bulletin of the American
Meteorological Society, Vol. 95, No 9, 2014, pp. 1431-1443.

Bonelli, S., ‘The first red list of Italian butterflies’, Insect Conservation and Diversity, Vol. 11, 2018, pp. 506–
5821.

Brown, I. and Castellazzi, M., ‘Scenario analysis for regional decision making on sustainable multifunctional land
uses’, Regional Environmental Change, Vol. 14, 2014, p. 15.

Brunetti, G., Šimůnek, J., and Piro, P., ‘A comprehensive numerical analysis of the hydraulic behavior of a
permeable pavement’, Journal of Hydrology, Vol. 540, 2016, pp. 1146–1161.
https://doi.org/10.1016/j.jhydrol.2016.07.030

Brunetti, G., Šimůnek, J., Turco, M., and Piro, P., ‘On the use of surrogate-based modeling for the numerical
analysis of low impact development techniques’, Journal of Hydrology, Vol. 548, 2017, pp. 263–277.
https://doi.org/10.1016/j.jhydrol.2017.03.013

Bruse, M., ENVI-Met Implementation of the Gas/Particle Dispersion and Deposition Model PDDM, 2017. Available
online: http://www.envi-met.net/documents/sources.PDF

California Department of Water Resources, Monitoring Networks and Identification of Data Gaps BMP, 2016, pp.
1-34.

Chen, H., Ooka, R., Huang, H., and Tsuchiya, T., ‘Study on mitigation measures for outdoor thermal environment
on present urban blocks in Tokyo using coupled simulation’, Building and Environment, Vol. 44, No 11,
2009, pp. 2290-2299.

Ciais, P., Dolman, A.J., Bombelli, A., Duren, R., Peregon, A., Rayner, P.J., Miller, C., Gobron, N., Kinderman, G.,
Marland, G., Gruber, N., Chevallier, F., Andres, R.J., Balsamo, G., Bopp, L., Bréon, F.-M., Broquet, G.,
Dargaville, R., Battin. T.J., Borges, A., Bovensmann, J.H., Buchwitz, M., Butler, J., Canadell, J.G., Cook,
R.B., DeFries, R., Engelen, R., Gurney, K.R., Heinze, C., Heimann, M., Held, A., Henry, M., Law, B.,

363
Luyssaert, S., Miller, J., Moriyama, T., Moulin, C., Myneni, R.B., Nussli, C., Obersteiner, M., Ojima, D.,
Pan, Y., Paris, J.-D., Piao, S.L., Poulter, B., Plummer, S., Quegan, S., Raymond, P, Reichstein, M., Rivier,
L, Sabine, C., Schimel, D., Tarasova, O., Valentini, R., Wang, R., van der Werf, G., Wickland, D., Williams,
M., and Zehner, C., ‘Current systematic carbon-cycle observations and the need for implementing a
policy-relevant carbon observing system’, Biogeosciences, Vol. 11, 2014, pp. 3547–3602. doi:
10.5194/bg-11-3547-2014

Clobert, J, Chanzy, A., Le Galliard, J.-F., Chabbi, A., Greiveldinger, L., Caquet, T., Loreau, M., Mougin, C., Pichot,
C., Roy, J. and Saint-André, L., ‘How to Integrate Experimental Research Approaches in Ecological and
Environmental Studies: AnaEE France as an Example’, Frontiers in Ecology and Evolution, Vol. 6, 2018,
p. 43. doi: 10.3389/fevo.2018.00043

Cohen, D.A., Setodji, C., Evenson, K.R., Ward, P., Lapham, S., Hillier, A., and McKenzie, T.L., ‘How much
observation is enough? Refining the administration of SOPARC’, Journal of Physical Activity and Health,
Vol. 8, No 8, 2011, pp. 1117-1123.

Corbane, C., Martino, P., Panagiotis, P., Aneta, F.J., Michele, M., Sergio, F., Marcello, S., Daniele, E., Gustavo,
N., and Thomas, K., ‘The grey-green divide: multi-temporal analysis of greenness across 10,000 urban
centres derived from the Global Human Settlement Layer (GHSL)’, International Journal of Digital Earth,
Vol. 13, No 1, 2020, pp. 101-118.

Creswell, J.W., ‘Mixed-method research: Introduction and application’, In: G.J. Cizek (Ed.) Handbook of
educational policy (pp. 455-472), Academic Press, 1999.

Csillag, F. and Boots., B., ‘A Framework for Statistical Infer-ential Decisions in Spatial Pattern Analysis’, Canadian
Geographer, Vol. 49, No 2, 2005, pp. 172–179.

Dickinson, J.L. and Bonney, J.R.E. (Eds.), Citizen science: Public participation in environmental research, Cornell
University Press, 2012.

Dickinson, J.L., Zuckerberg, B., and Bonter, D.N., ‘Citizen science as an ecological research tool: challenges and
benefits’, Annual review of ecology, evolution, and systematics, Vol. 41, 2010, pp. 149-172.

Dushkova, D. and Ignatieva, M., ‘New trends in urban environmental health research: from geography of diseases
to therapeutic landscapes and healing gardens’, Geography, Environment, Sustainability, Vol. 13, No 1,
2020, pp. 159-171. https://doi.org/10.24057/2071-9388-2019-99

Dushkova, D. and Dagmar, H., ‘Not Simply Green: Nature-Based Solutions as a Concept and Practical Approach
for Sustainability Studies and Planning Agendas in Cities’, Land, Vol. 9, 2020, p. 19.
10.3390/land9010019.

European Space Agency (ESA), Newcomers Earth Observation Guide, 2019. Available at:
https://business.esa.int/newcomers-earth-observation-guide.

European Parliament and of the Council of the European Union, ‘Directive 2007/2/EC of the European Parliament
and of the Council of 14 March 2007 establishing an Infrastructure for Spatial Information in the European
Community (INSPIRE)’, Official Journal of the European Union L 108, 2007, p. 1–14.

European Parliament and of the Council of the European Union, ‘Regulation (EU) 2016/679 of the European
Parliament and of the Council of 27 April 2016 on the protection of natural persons with regard to the
processing of personal data and on the free movement of such data, and repealing Directive 95/46/EC’,
Official Journal of the European Union L 119, 2016, pp. 1-88.

Statistical Office of the European Union (Eurostat), Glossary, 2014. https://ec.europa.eu/eurostat/statistics-


explained/index.php/Glossary:Baseline_study

Statistical Office of the European Union (Eurostat), Statistical regions in the European Union and partner
countries: NUTS and statistical regions 2021, Publications Office of the European Union, Luxembourg,
2020. https://ec.europa.eu/eurostat/documents/3859598/10967554/KS-GQ-20-092-EN-
N.pdf/9d57ae79-3ee7-3c14-da3e-34726da385cf

Frantzeskaki, N., McPhearson, T., Collier, M. J., Kendal, D., Bulkeley, H., Dumitru, A., Walsh, C., Noble, K., van
Wyk, E., Ordóñez, C., Oke, C., and Pintér, L., ‘Nature-based solutions for urban climate change
adaptation: linking science, policy, and practice communities for evidence-based decision-making’,
BioScience, Vol. 69, No 6, 2019, pp. 455-466.

Frumkin, H., Bratman, G.N., Breslow, S.J., Cochran, B., Kahn Jr., P.H., Lawler, J.J., Levin, P.S., Tandon, P.S.,
Varanasi, U., Wolf, K.L. and Wood, S.A., ‘Nature contact and human health: A research agenda’,
Environmental health perspectives, Vol. 125, No 7, 2017, p. 075001.

Gardiner, M.M., Allee, L.L., Brown, P.M., Losey, J.E., Roy, H.E., and Smyth, R.R., ‘Lessons from lady beetles:
accuracy of monitoring data from US and UK citizen‐science programs’, Frontiers in Ecology and the
Environment, Vol. 10, No 9, 2012, pp. 471-476.

364
Gittleman, M., Farmer, C.J., Kremer, P., and McPhearson, T., ‘Estimating stormwater runoff for community
gardens in New York City’, Urban ecosystems, Vol. 20, No 1, 2017, pp. 129-139.

Government of South Australia, ‘Chapter 15: Modelling Process and tools’, Water Sensitive Urban Design
Technical Manual for the Greater Adelaide Region, Government of South Australia, Adelaide, 2010, p. 48.

Grizzetti, B., Bouraoui, F. and Aloe, A., ‘Changes of nitrogen and phosphorus loads to European seas’, Global
Change Biology, Vol. 18, 2012, pp. 769-782.

Groenewegen, P.P., van den Berg, A.E., de Vries, S. and Verheij, R.A., ‘Vitamin G: effects of green space on
health, well-being, and social safety’, BMC Public Health, Vol. 6, 2006, p. 149.

Gruiz, K., Meggyes, T., and Fenyvesi, E., Engineering Tools for Environmental Risk Management: 3. Site
Assessment and Monitoring Tools, CRC Press, 2017.

Grützner, R., ‘Environmental modeling and simulation - applications and future requirements’, Environmental
Software Systems, Springer Science+Business Media, Dordrecht, 1996.

Guida, R.J., Swanson, T.L., Remo, J.W., and Kiss, T., ‘Strategic floodplain reconnection for the Lower Tisza River,
Hungary: opportunities for flood-height reduction and floodplain-wetland reconnection’, Journal of
Hydrology, Vol. 521, 2015, pp. 274-285.

Guo, B., Wang, Z., Yu, Z., Wang, Y., Yen, N., Huang, R., and Zhou, X., ‘Mobile Crowd Sensing and Computing:
The Review of an Emerging Human-Powered Sensing Paradigm’, ACM Computing Surveys, Vol. 48, 2015.

Haase D., Haase A., Kabisch N., Kabischb S., and Rink D., ‘Actors and factors in land-use simulation: The
challenge of urban shrinkage’, Environmental Modelling and Software, Vol. 35, 2012, pp. 92-103.

Hamad R., Balzter H., and Kolo K., ‘Predicting land use/land cover changes using a CA-Markov model under two
different scenarios’, Sustainable, Vol. 10, 2018, p. 3421.

Hampton, S.E., Strasser, C.A., Tewksbury, J.J., Gram, W.K., Budden, A.E., Batcheller, A.L., Duke, C.S., and
Porter, J.H., ‘Big data and the future of ecology’, Frontiers in Ecology and the Environment, Vol. 11, No
3, 2013, pp. 156-162.

Hansen, R., Rolf, W., Santos, A., Luz, A.C., Száraz, L., Tosics, I., Vierikko, K., Rall, E., Davies, C., and Pauleit,
S., Advanced Urban Green Infrastructure Planning and Implementation - Innovative Approaches and
Strategies from European Cities, Deliverable 5.2. Technical Report of the Green Surge Project, 2016.

Hartig T., Mitchell R., de Vries S., and Frumkin H., ‘Nature and health’, Annual Review of Public Health, Vol. 35,
2014, pp. 207–228.

Hecker, S., Haklay, M., Bowser, A., Makuch, Z., Vogel, J., and Bonn, A., Citizen Science: Innovation in Open
Science, Society and Policy, UCL Press, London, 2018.

Herold, M., Couclelis, H., and Clarke, K.C., ‘The role of spatial metrics in the analysis and modeling of urban land
use change’, Computers, environment and urban systems, Vol. 29, No 4, 2005, pp. 369-399.

Hirabayashi, S., Kroll, C.N., and Nowak, D.J., ‘Development of a distributed air pollutant dry deposition modeling
framework’, Environmental Pollution, Vol. 171, 2012, pp. 9–17.

Hodson, S., Collins, S., Genova, F., Harrower, N., Jones, S., Laaksonen, L., Mietchen, D., Petrauskaité, R. and
Wittenburg, P., Turning FAIR into reality: Final report and action plan from the European Commission
expert group on FAIR data, Publications Office of the European Union, Luxembourg, 2018.

Houghton, J., Townshend, J., Dawson, K., Mason, P., Zillman, J., and Simmons, A., ‘The GCOS at 20 years: the
origin, achievement and future development of the Global Climate Observing System’, Weather, Vol. 67,
2012, pp. 227-235.

Iacob, O., Rowan, J.S., Brown, I.M., and Ellis, C., ‘Evaluating wider benefits of natural flood management
strategies: An ecosystem-based adaptation perspective’, Hydrology Research, Vol. 45, No 6, 2014, pp.
774-787.

Ilieva, R.T. and McPhearson, T., ‘Social-media data for urban sustainability’, Nature Sustainability, Vol. 1, No 10,
2018, pp. 553-565.

INSPIRE, INSPIRE Knowledge Base, 2019. Available from https://inspire.ec.europa.eu. Accessed 2019, July 29.

International Federation of Red Cross and Red Crescent Societies, Baseline Basics, 2013.

International Organization for Standardization (ISO), ISO 19101-1:2014, Geographic information – Reference
model – Part 1: Fundamentals, International Organization for Standardization, Geneva, 2014.

International Organization for Standardization (ISO), ISO/TC 211 Geographic information/Geomatics, 2019.

365
Fernandez, F.J., Alvarez-Velazquez, L.J., Garcia-Chan, N., Martinez, A., and Velazquez-Mendez, M.E., ‘Optimal
location of green zones in metropolitan areas to control the urban heat island’, Journal of Computational
and Applied Mathematics, Vol. 289, 2015, pp. 412–425.

Filibeck, G., Petrella, P., and Cornelini, P., ‘All ecosystems look messy, but some more so than others: A case-
study on the management and acceptance of Mediterranean urban grasslands’, Urban Forestry and Urban
Greening, Vol. 15, 2016, pp. 32–39.

Florczyk, A.J., Corbane, C., Ehrlich, D., Carneiro Freire, S.M., Kemper, T., Maffenini, L., Melchiorri, M., Pesaresi,
M., Politis, P., Schiavina, M., Sabo, F., and Zanchetta, L., GHSL data package 2019, Publications Office
of the European Union, Luxembourg, 2019.

Hawxwell, T., Mok, S., Mačiulyte, E., Sautter, J., Theobald, J.A., Dobrokhotova, E., and Suska, P., Municipal
Governance Guidelines, Urban Nature Labs (UNaLab) Deliverable D6.2, 2018.

Kabisch, N. and Haase, D., ‘Urban nature benefits – Opportunities for improvement of health and well-being in
times of global change’, Newsletter on Housing and Health, Vol. 29, 2018, pp. 1-11.

Kabisch, N., Stadler, J., Korn H., and Bonn, A., Nature-based solutions to climate change mitigation and
adaptation in urban areas, Federal Agency for Nature Conservation, Bonn, 2016.

Kabisch, N., van den Bosch, M., and Lafortezza, R., ‘The health benefits of nature-based solutions to urbanization
challenges for children and the elderly – A systematic review’, Environmental Research, Vol. 159, 2017,
pp. 362-373.

Kawulich, B., ‘Collecting data through observation’, Doing Social Research: A global context, McGraw-Hill,
Berkshire, 2012.

Krebs, G., Kokkonen, T., Valtanen, M., Setälä, H., and Koivusalo, H., ’Spatial resolution considerations for urban
hydrological modelling’, Journal of Hydrology, Vol. 512, 2014, pp. 482-497.

Kumar, P., Druckman, A., Gallagher, J., Gatersleben, B., Allison, S., Eisenman, T.S., Hoang, U., Hama, S., Tiwari,
A., Sharma, A., Abhijth, K.V., Aslakha, D., McNabola, A., Astell-Burt, T., Feng, X., Skeldon, A.C., de
Lusignan, S., and Morawska, L., ‘The nexus between air pollution, green infrastructure and human
health’, Environment international, Vol. 133 A, 2019, p. 105181.

Langevin, C.D., Hughes, J.D., Banta, E.R., Niswonger, R.G., Panday, S., and Provost, A.M., Documentation for
the MODFLOW 6 groundwater flow model (No. 6-A55), US Geological Survey, 2017.

Lein, J.K., Environmental Sensing, Springer, New York, 2012.

Leopa, S. and Elisei, P., Methodology on spatial analysis in front-runner and follower cities, Deliverable No. 2.1,
proGIreg. Horizon 2020 Grant Agreement No 776528, European Commission, 2020, 61 pp.

Leopa, S.; Elisei, P., Knappe, D., Koeneke, H., and Nohra, G.C., Spatial Analysis in Front-Runner and Follower
Cities, Deliverable No. 2.2, proGIreg. Horizon 2020 Grant Agreement No. 776528, European Commission,
2019.

Li, T., Guo, S., An, D., and Nametso, M., ‘Study on water and salt balance of plateau salt marsh wetland based
on time-space watershed analysis’, Ecological Engineering, Vol. 138, 2019, pp. 160-170.

Linden, M. and Sheehy, N., ‘Comparison of a Verbal Questionnaire and Map in Eliciting Environmental
Perceptions’, Environment and Behavior, Vol. 36, No 1, 2004, pp. 32–40.

Litwin, L. and Rossa, M., Geoinformation Metadata in INSPIRE and SDI: Understanding. Editing. Publishing,
Springer Science and Business Media, 2011.

Luettich, R.A., Jr., Westerink, J.J., and Scheffner, N.W., ADCIRC: An Advanced Three-Dimensional Circulation
Model for Shelves, Coasts, and Estuaries, Report 1: Theory and Methodology of ADCIRC-2DDI and
ADCIRC-3DL, Dredging Research Program Technical Report DRP-92-6, U.S. Army Engineers Waterways
Experiment Station, Vicksburg, 1992.

Lukyanenko, R., Parsons, J., and Wiersma, Y.F., ‘Emerging problems of data quality in citizen science’,
Conservation Biology, Vol. 30, 2016, pp. 447-449.

Mahmood, A.A. and Elektorowicz, M., ‘A review of discrete element method research on particulate systems’, In
IOP Conference Series: Materials Science and Engineering, Vol. 136, No 1, 2016, p. 012034.

Maldonado, G. and Greenland, S., ‘Estimating causal effects’, International Journal of Epidemiology, Vol. 31, No
2, 2002, pp. 422–429.

Marcel, M.R., Stadler, J., Korn, H., Irvine, K., and Bonn, A. (Eds.), Biodiversity and Health in the Face of Climate
Change, Springer International Publishing, Cham, 2019.

Mason, D.C., Horritt, M.S., Hunter, N.M., and Bates, P.D., ‘Use of fused airborne scanning laser altimetry and
digital map data for urban flood modelling’, Hydrological Processes, Vol. 21, No 11, 2007, pp. 1436-1447.

366
Matos, P., Vieira, J., Rocha, B., Branquinho, C., and Pinho, P., ‘Modeling the provision of air-quality regulation
ecosystem service provided by urban green spaces using lichens as ecological indicators’, Science of The
Total Environment, Vol. 665, 2019, pp. 521-530.

McDonald, A.G., Bealey, W.J., Fowler, D., Dragosits, U., Skiba, U., Smith, R.I., Donovan, R.G., Brett, H.E., Hewitt,
C.N., and Nemitz, E., ‘Quantifying the effect of urban tree planting on concentrations and depositions of
PM10 in two UK conurbations’, Atmospheric Environment, Vol. 41. No 38, 2007, pp. 8455-8467.

McKenzie, T.L., Cohen, D.A., Sehgal, A., Williamson, S., and Golinelli, D., ‘System for Observing Play and
Recreation in Communities (SOPARC): Reliability and Feasibility Measures’, Journal of Physical Activity
and Health, Vol. 3, Suppl. 1, 2006, pp. S208–S222.

Mechoso, C.R. and Arakawa, A., ‘NUMERICAL MODELS: General Circulation Models’, In G.R. North, Pyle, J.A., and
Zhang, F. (Eds.), Encyclopedia of Atmospheric Sciences (2nd ed.), Academic Press, 2015, pp. 153-160.

Mohareb, E. and Kennedy, C., ‘Gross direct and embodied carbon sinks for urban inventories’, Journal of Industrial
Ecology, Vol. 16, No 3, 2012, pp. 302-316.

Montgomery, D.C., Design and analysis of experiments, John Wiley & Sons, New York, 2008.

Morales-Torres, A., Escuder-Bueno, I., Andrés-Doménech, I., and Perales-Momparler, S., ‘Decision Support Tool
for energy-efficient, sustainable and integrated urban stormwater management’, Environmental
Modelling and Software, Vol. 84, 2016, pp. 518-528.

Nelson, E., Mendoza, G., Regetz, J., Polasky, S., Tallis, H., Cameron, D. R., Chan, K.M.A., Daily, G.C., Goldstein,
J., Kareiva, P.M., Lonsdorf, E., Naidoo, R., Ricketts, T.H., and Shaw, M.R., ‘Modeling multiple ecosystem
services, biodiversity conservation, commodity production, and tradeoffs at landscape scales’, Frontiers
in Ecology and the Environment, Vol. 7, No 1, 2009, pp. 4–11.

Nelson E.J. and Daily G.C., ‘Modelling ecosystem services in terrestrial systems’, F1000 Biology Reports, Vol. 2,
2010.

Nieto, A., Roberts, S.P.M., Kemp, J., Rasmont, P., Kuhlmann, M., García Criado, M., Biesmeijer, J.C., Bogusch,
P., Dathe, H.H., De la Rúa, P., and De Meulemeester, T., European Red List of Bees, Publication Office of
the European Union, Luxembourg, 2014.

Nijhuis, S., Zlatanova, S., Dias, E., van der Hoeven, F., and van der Spek, S. (Eds.), ‘Geo-Design Advances in
bridging geo-information technology, urban planning and landscape architecture’, Research in Urbanism,
Vol. 4, TU Delft Open, Delft, 2014.

Nowak, D.J., Hirabayashi, S., Bodine, A., and Greenfield, E., ‘Tree and forest effects on air quality and human
health in the United States’, Environmental Pollution, Vol. 193, 2014, pp. 119–129.

Open Geospatial Consortium (OGC), OGC Sensor Observation Service Interface Standard (v2.0), 2012.

Open Geospatial Consortium (OGC), Open Geospatial Consortium, 2019. Available from
http://www.opengeospatial.org/. Accessed 2019, July 18.

Orgiazzi, A., Ballabio, C., Panagos, P., Jones, A., and Fernández‐Ugalde, O., ‘LUCAS Soil, the largest expandable
soil dataset for Europe: a review’, European Journal of Soil Science, Vol. 69, No 1, 2018, pp. 140-153.

Pain, H., ‘A literature review to evaluate the choice and use of visual methods’, International Journal of Qualitative
Methods, Vol. 11, 2012, pp. 303–319.

Pastorello, G., Papale, D., Chu, H., Trotta, C., Agarwal, D., Canfora, E., Baldocchi, D., and Torn, M., ‘A new data
set to keep a sharper eye on land-air exchanges’, Eos, Transactions American Geophysical Union, Vol.
98, No 8, 2017.

Pepper, I.L., Brusseau, M.L., and Artiola, J., Environmental monitoring and characterization, Elsevier Science &
Technology, 2004.

Pesaresi, M., ‘Principles and Applications of the Global Human Settlement Layer’, IGARSS 2018 - 2018 IEEE
International Geoscience and Remote Sensing Symposium, Valencia, 2018, pp. 2047–2050.

Pocock, M.J.O., Chapman, D.S., Sheppard, L.J., and Roy, H.E., A Strategic Framework to Support the
Implementation of Citizen Science for Environmental Monitoring, Final report to SEPA, Centre for Ecology
and Hydrology, Wallingford, 2014.

Pollard, E. and Yates, T.J., Monitoring butterflies for ecology and conservation, Chapman & Hall, New York. 1993.

Ponto, J., ‘Understanding and Evaluating Survey Research’, Journal of The Advanced Practitioner in Oncology,
Vol. 6, No 2, 2015, pp. 168–171.

Pregnolato, M., Ford, A., Robson, C., Glenis, V., Barr, S., and Dawson, R., ‘Assessing urban strategies for reducing
the impacts of extreme weather on infrastructure networks’, Royal Society Open Science, Vol. 3, No 5,
2016.

367
Raymond, C.M., Pam, B., Breil, M., Nita, M.R., Kabisch, N., de Bel, M., Enzi, V., Frantzeskaki, N., Geneletti, D.,
Cardinaletti, M., Lovinger, L., Basnou, C., Monteiro, A., Robrecht, H., Sgrigna, G., Munari, L., and
Calfapietra, C., An Impact Evaluation Framework to Support Planning and Evaluation of Nature-based
Solutions Projects. Report prepared by the EKLIPSE Expert Working Group on Nature-based Solutions to
Promote Climate Resilience in Urban Areas, Centre for Ecology and Hydrology, Wallingford, 2017.

Reyers, B., Stafford-Smith, M., Erbs, K.-H., Scholes, R.J., and Selomane, O., ‘Essential Variables help to focus
Sustainable Development Goals monitoring’, Current Opinion in Environmental Sustainability, Vol. 26,
2017, pp. 97–105.

Rosenbaum, P.R., Design of observational studies, Springer, New York, 2010.

Rossman, L.A., Storm Water Management Model User’s Manual Version 5.1, United States Environmental
Protection Agency, Cincinnati, 2015.

Sahukhal R. and Bajracharya T.R., ‘Modeling water resources under competing demands for sustainable
development: A case study of Kaligandaki Gorge Hydropower Project in Nepal’, Water Science and
Engineering, Vol. 12, No 1, 2019, pp. 19-26.

Sang, N. (Ed.), Modelling Nature-Based Solutions: Integrating Computational and Participatory Scenario
Modelling for Environmental Management and Planning, Cambridge University Press, Cambridge, 2020.

Schebella, M. F., Weber, D., Schultz, L., and Weinstein, P., ‘The Nature of Reality: Human Stress Recovery during
Exposure to Biodiverse, Multisensory Virtual Environments’, International Journal of Environmental
Research and Public Health, Vol. 17, No 1, 2020, p. 56.

Schiavina, M., Melchiorri, M., Corbane, C., Florczyk, A.J., Freire, S., Pesaresi, M., and Kemper, T., ‘Multi-Scale
Estimation of Land Use Efficiency (SDG 11.3.1) across 25 Years Using Global Open and Free Data’,
Sustainability, Vol. 11, No 20, 2019, p.5674.

Scholes, R.J., Reyers, B., Biggs, R., Spierenburg, M.J., and Duriappah, A., ‘Multi-scale and cross-scale
assessments of social–ecological systems and their ecosystem services’, Current Opinion in
Environmental Sustainability, Vol. 5, No 1, 2013, pp. 16-25.

Schwarz, N., Kahlenberg, D., Haase, D., and Seppelt, R., ‘ABMland - A tool for collaborative agent-based model
development on urban land use change’, Journal of Artificial Societies and Social Simulation, Vol. 15, No
2, 2012, p. 8.

Sekulova, F., Anguelovski, I., Argüelles, L., and Conill, J., ‘A ‘fertile soil’ for sustainability-related community
initiatives: A new analytical framework’, Environment and Planning A, Vol. 49, No 10, 2017, pp. 2362–
2382.

Serrano Sanz, F., Holocher-Ertl, T., Kieslinger, B., Sanz Garcia, F., and Silva, C.G., White Paper on citizen science
for Europe, 2014. Available from
https://ec.europa.eu/futurium/en/system/files/ged/socientize_white_paper_on_citizen_science.pdf

Seyfang, G. and Longhurst, N., ‘What influences the diffusion of grassroots innovations for sustainability?
Investigating community currency niches’, Technology Analysis and Strategic Management, Vol. 28, No
1, 2016, pp. 1–23.

Simon, H., Fallmann, J., Kropp, T., Tost, H., and Bruse, M., ‘Urban Trees and Their Impact on Local Ozone
Concentration — A Microclimate Modeling Study’, Atmosphere, Vol. 10, No 3, 2019, p. 154.

Smith, K.R. and Roebber, P.J., ‘Green roof mitigation potential for a proxy future climate scenario in Chicago,
Illinois’, Journal of Applied Meteorology and Climatology, Vol. 50, No 3, 2011, pp. 507-522.

Stark, J., Plancke, Y., Ides, S., Meire, P., and Temmerman, S., ‘Coastal flood protection by a combined nature-
based and engineering approach: Modeling the effects of marsh geometry and surrounding dikes’,
Estuarine, Coastal and Shelf Science, Vol. 175, 2016, pp. 34-45.

Steinitz, C., ‘Beginnings of Geodesign: a personal historical perspective’, Research in Urbanism Series, Vol. 4,
2016, pp. 9-22.

Sullivan, L.M., Weinberg, J., and Keaney Jr., J.F., ’Common statistical pitfalls in basic science research’, Journal
of the American Heart Association, Vol. 5, No 10, 2016, p. e004142.

Sun, C.Y., Lee, K.P., Lin, T.P., and Lee, S.H., ‘Vegetation as a Material of Roof and City to cool down the
temperature’, Advanced Materials Research, Vol. 461, 2012, pp. 552-556.

Sun, L., Wang, W., Gao, F., Cao, C., and Tang, J., ‘Distribution and Influencing Factors of Plankton in an Artificial
Lagoon in Ningbo’, Asian Journal of Ecotoxicology, Vol. 4, 2018, pp. 60-67

Surussavadee, C., ‘Evaluation of WRF near-surface wind simulations in tropics employing different planetary
boundary layer schemes’, In 8th International Renewable Energy Congress (IREC), Amman, 2017, pp.
1-4.

368
ten Brink, P., Mutafoglu, K., Schweitzer, J.P., Kettunen, M., Twigger-Ross, C., Baker, J., Kuipers, Y., Emonts, M.,
Tyrväinen, L., Hujala, T., and Ojala, A., The health and social benefits of nature and biodiversity
protection, Institute for European Environmental Policy, London/Brussels, 2016.

Thompson, J.R., Iravani, H., Clilverd, H.M., Sayer, C.D., Heppell, C.M., and Axmacher, J.C., ‘Simulation of the
hydrological impacts of climate change on a restored floodplain’, Hydrological Sciences Journal, Vol. 62,
2017, pp. 2482-2510.

Underwood, E., Darwin, G., and Gerritsen, E., Pollinator initiatives in EU Member States: Success factors and
gaps. Report for European Commission under contract for provision of technical support related to Target
2 of the EU Biodiversity Strategy to 2020 – maintaining and restoring ecosystems and their services,
Institute for European Environmental Policy, Brussels, 2017.

Vercelloni, J., Clifford, S., Caley, M.J., Pearse, A.R., Brown, R., James, A., Christensen, B., Bednarz, T., Anthony,
K., González-Rivero, M., and Mengersen, K., ‘Using virtual reality to estimate aesthetic values of coral
reefs’, Royal Society Open Science, Vol. 5, No 4, 2018, p.172226.

Vranic, S., Kalas, M. Leo, L.S., Bertini, F., Sabbatini, T., Neupane, B., Debele, E., and Barisani F., GeoIKP
Deployment and Maintenance Report, Deliverable No. 7.9, OPERANDUM, Horizon 2020, Grant Agreement
No. 776848, European Commission, 2019.

Vuik, V., Jonkman, S.N., Borsje, B.W., and Suzuki, T., ‘Nature-based flood protection: The efficiency of vegetated
foreshores for reducing wave loads on coastal dikes’, Coastal engineering, Vol. 116, 2016, pp. 42-56.

Wamsley, T.V., Cialone, M.A., Smith, J.M., Atkinson, J.H., and Rosati, J.D., ‘The potential of wetlands in reducing
storm surge’, Ocean Engineering, Vol. 37, 2010, pp. 59–68.

Weathers, K.C., Hanson, P.C., Arzberger, P., Brentrup, J., Brookes, J., Carey, C.C., Gaiser, E., Hamilton, D.P.,
Hong, G.S., Ibelings, B., and Istvánovics, V., ‘The global lake ecological observatory network (GLEON):
The evolution of grassroots network science’, Limnology and Oceanography Bulletin, Vol. 22, No 3, 2013,
pp.71-73.

World Health Organization (WHO), Urban Green Space and Health: Intervention Impacts and Effectiveness.
Meeting report. Bonn, Germany, 20–21 September 2016, WHO Regional Office for Europe, Copenhagen,
2017.

Wilkinson, M.D., Dumontier, M., Aalbersberg, I.J., Appleton, G., Axton, M., Baak, A., Blomberg, N., Boiten, J.W.,
da Silva Santos, L.B., Bourne, P.E., and Bouwman, J., ‘The FAIR Guiding Principles for scientific data
management and stewardship’, Scientific Data, Vol. 3, No 1, 2016, pp. 1-9.

Williams, A., Therapeutic Landscapes, Routledge, London, 2017.

Wolfram, M., ‘Cities shaping grassroots niches for sustainability transitions: Conceptual reflections and an
exploratory case study’, Journal of Cleaner Production, Vol. 173, 2018, pp. 11–23.

Wood, C.J., Pretty, J., and Griffin, M., ‘A case-control study of the health and well-being benefits of allotment
gardening’, Journal of Public Health, Vol. 38, 2016, pp. e336-e344.

World Bank, Implementing nature-based flood protection: Principles and implementation guidance, World Bank,
Washington, DC, 2017.

World Meteorological Organization (WMO), Essential Climate Variables, 2020. Available from
https://public.wmo.int/en/programmes/global-climate-observing-system/essential-climate-variables.

Woroniecki, S., ‘Enabling Environments? Examining Social Co-Benefits of Ecosystem-Based Adaptation to Climate
Change in Sri Lanka’, Sustainability, Vol. 11, No 3, 2019, p. 772.

Yuan, Z. Y., Jiao, F., Shi, X. R., Sardans, J., Maestre, F. T., Delgado-Baquerizo, M., Reich, P. B., and Peñuelas,
J., ‘Experimental and observational studies find contrasting responses of soil nutrients to climate change’,
eLife, Vol. 6, 2017, e23255.

Zulian, G., Polce, C., and Maes, J., ‘ESTIMAP: a GIS-based model to map ecosystem services in the European
union’, Annali di Botanica, Vol. 4, 2014, pp. 1-7.

369
Getting in touch with the EU
IN PERSON
All over the European Union there are hundreds of Europe Direct information centres.
You can find the address of the centre nearest you at:
https://europa.eu/european-union/contact_en

ON THE PHONE OR BY EMAIL


Europe Direct is a service that answers your questions about the European Union.
You can contact this service:
– by freephone: 00 800 6 7 8 9 10 11 (certain operators may charge for these calls),
– at the following standard number: +32 22999696, or
– by email via: https://europa.eu/european-union/contact_en

Finding information about the EU


ONLINE
Information about the European Union in all the official languages of the EU is available on the Europa
website at: https://europa.eu/european-union/index_en

EU PUBLICATIONS
You can download or order free and priced EU publications from:
https://op.europa.eu/en/publications. Multiple copies of free publications may be obtained by contacting
Europe Direct or your local information centre (see https://europa.eu/european-union/contact_en)

EU LAW AND RELATED DOCUMENTS


For access to legal information from the EU, including all EU law since 1952 in all the official language
versions, go to EUR-Lex at: http://eur-lex.europa.eu

OPEN DATA FROM THE EU


The EU Open Data Portal (http://data.europa.eu/euodp/en) provides access to datasets from the EU. Data
can be downloaded and reused for free, for both commercial and non-commercial purposes.
The Handbook aims to provide decision-makers with a comprehensive
NBS impact assessment framework, and a robust set of indicators and
methodologies to assess impacts of nature-based solutions across 12
societal challenge areas: Climate Resilience; Water Management; Natural
and Climate Hazards; Green Space Management; Biodiversity; Air Quality;
Place Regeneration; Knowledge and Social Capacity Building for Sustainable
Urban Transformation; Participatory Planning and Governance; Social Justice
and Social Cohesion; Health and Well-being; New Economic Opportunities
and Green Jobs.

Indicators have been developed collaboratively by representatives of 17


individual EU-funded NBS projects and collaborating institutions such
as the EEA and JRC, as part of the European Taskforce for NBS Impact
Assessment, with the four-fold objective of: serving as a reference for
relevant EU policies and activities; orient urban practitioners in developing
robust impact evaluation frameworks for nature-based solutions at different
scales; expand upon the pioneering work of the EKLIPSE framework by
providing a comprehensive set of indicators and methodologies; and build
the European evidence base regarding NBS impacts. They reflect the state
of the art in current scientific research on impacts of nature-based solutions
and valid and standardized methods of assessment, as well as the state of
play in urban implementation of evaluation frameworks.

Studies and reports

You might also like