You are on page 1of 96

SpringerBriefs in

Architectural Design and Technology


Pia Fricker · Toni Kotnik

Patterns of Interaction
Computational Design
Across Scales
SpringerBriefs in Architectural Design
and Technology

Series Editor
Thomas Schröpfer, Architecture and Sustainable Design, Singapore
University of Technology and Design, Singapore, Singapore
Indexed by SCOPUS
Understanding the complex relationship between design and technology is
increasingly critical to the field of Architecture. The Springer Briefs in
Architectural Design and Technology series provides accessible and
comprehensive guides for all aspects of current architectural design relating to
advances in technology including material science, material technology, structure
and form, environmental strategies, building performance and energy, computer
simulation and modeling, digital fabrication, and advanced building processes. The
series features leading international experts from academia and practice who
provide in-depth knowledge on all aspects of integrating architectural design
with technical and environmental building solutions towards the challenges of a
better world. Provocative and inspirational, each volume in the Series aims to
stimulate theoretical and creative advances and question the outcome of technical
innovations as well as the far-reaching social, cultural, and environmental
challenges that present themselves to architectural design today. Each brief asks
why things are as they are, traces the latest trends and provides penetrating,
insightful and in-depth views of current topics of architectural design. Springer
Briefs in Architectural Design and Technology provides must-have, cutting-edge
content that becomes an essential reference for academics, practitioners, and
students of Architecture worldwide.
Pia Fricker · Toni Kotnik

Patterns of Interaction
Computational Design Across Scales
Pia Fricker Toni Kotnik
Aalto University Aalto University
Espoo, Finland Espoo, Finland

ISSN 2199-580X ISSN 2199-5818 (electronic)


SpringerBriefs in Architectural Design and Technology
ISBN 978-981-19-9082-3 ISBN 978-981-19-9083-0 (eBook)
https://doi.org/10.1007/978-981-19-9083-0

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Foreword: Perpetual Boundary Judgments

Patterns of Interaction addresses timely topics and concepts that permeate contem-
porary discourses. These link together urban design, architecture, and landscape
architecture from an environmental perspective, and include urbanization, ground,
topology, performance and implicitly also environment, in the greater sense, as a set
of complex dynamic relations. The authors outline a particular approach to design
involving a methodological framework that combines aspects of design thinking,
systems thinking, and design computing. Tackling such multifaceted topics inevitably
necessitates the making of frequent boundary judgments, i.e., timely and spontaneous
decisions as to what is to be included or excluded from the approach and from the
design agenda as the approach is developed.
Werner Ulrich developed Critical Systems Heuristics [1], a key concept of critical
systems thinking, as a conceptual framework for boundary critique [2]. In general,
professional propositions require choices regarding which facts (observations) and
norms (valuation standards) are considered relevant and are included, as well as
those, which are less relevant and can be left out.
However, every aspect that requires boundary judgments can also be subject to
change. This already begins at a primary conceptual level when insights and discourse
change, as, for instance, in the case with what constitutes an environment [3], and
furthermore what role politics [4], practices [5] or experiences [6] can play in offset-
ting previous understandings or opening new inroads. When defining or selecting a
conceptual approach, for instance, in the notion of performance, it is necessary to
select related items, relations, and dynamics. Performance requires agency, which
indicates the capacity with which something is able to act in the world. Actor-Network
Theory positioned agency as a non-human trait, thereby making it possible to think
of all human and non-human entities as actors or actants possessing agency [7]. This
entails an understanding of the performative capacity of the items and therefore also
of their dynamic relations to other items in their setting [8]. Given that these dynamic
relations form networks, it becomes clear that boundary judgments are ephemeral
and might change in any process of environmental transformation, e.g., by way of
urbanization and construction.

v
vi Foreword: Perpetual Boundary Judgments

Given today’s prolific pursuit and development of discourse, primary and


secondary concepts are frequently reframed. This makes it very challenging to
address, as this book does, questions of urbanization, ground, topology, performance,
and related issues of figure–ground, grounding, operativity (presumably from a prag-
matist perspective), patterns, organized matter, etc. And in this context clearly no
boundary judgment is without immediate and substantial consequence.
The same is true for boundary judgments related to the methodological approach.
Design thinking and systems thinking deliver useful inroads in terms of framing
design problems and solution spaces. Yet again, there exist a wide array of different
approaches. The same applies to design computing.
Today, Ian McHargs’ work is a frequent key reference, not only in terms of
approach, but also in terms of establishing a specific methodological use of data and
thereby initiating Geographic Information Systems [9]. Still, the question remains,
as to which aspects, dynamics, and scale ranges should be selected and focused on.
Pim Martens stated, “a new research paradigm is needed that is better able to reflect
the complexity and the multi-dimensional character of sustainable development.
The new paradigm … must be able to encompass different magnitudes of scale (of
time, space, and function), multiple balances (dynamics), multiple actors (interests)
and multiple failures (systemic faults)” [10]. To incorporate this in a computational
approach presents numerous challenges but entails, in the first instance, the ques-
tion whether to aim for an all-inclusive type of world model or an approach that is
based on models that are custom assembled and tailored to a context-specific design
task. Furthermore, data-integrated approaches to trans-scalar urban and architectural
design are attracting growing interest and attention and require careful and thorough
examination [11–13] .
That this is not just theoretical deliberation becomes clear when examining
approaches related to the one revealed in this book, overlapping with and differing
from Landscape Urbanism [14] or various concepts of ground [15–17].
Given all these challenges, books like Patterns of Interaction constitute impor-
tant contributions to the development of design approaches that link urban design,
architecture, and landscape architecture in order to provide more sound and
environmentally and ecologically positive results. Great insight may be gained into
their understanding when such works disclose the boundary judgments that were
continually made to progress in their development.

Vienna, Austria Michael Hensel

References

1. Ulrich W (1983) Critical heuristics of social planning: A new approach to practical philosophy.
Wiley, London
2. Ulrich W (1996) A primer to critical systems heuristics for action researchers. Centre for
Systems Studies, University of Hull
Foreword: Perpetual Boundary Judgments vii

3. Benson ES (2020) Surroundings—A history of environments and environmentalisms. The


University of Chicago Press, Chicago
4. Latour B (2018) Down to earth: politics in the new climatic regime. Polity Press, Cambridge
5. Hensel M (2019) The rights to ground: integrating human and non-human perspectives in an
inclusive approach to sustainability. Sustainable Development, 27, 245–251
6. Ingold T (2000) The perception of the environment—essays on livelihood, dwelling and skill.
Routledge: London
7. Latour B (2005) Reassembling the social: An introduction to actor-network Theory. Oxford
University Press, London
8. Hensel M (2013) Performance-oriented architecture—rethinking architectural design and the
built environment. Wiley: London
9. McHarg IL (1969) Design with nature. Doubleday, Garden City
10. Martins P (2006) ‘Sustainability: science or fiction?’. Sustainability: Sci Pract Policy 1(2):
36–41
11. Hensel M, Sørensen SS (2019) Performance-oriented architecture and urban design—relating
information-based design and systems-thinking in architecture. FORMAkademisk 12(2):1–17
12. Sunguroğlu Hensel D, Tyc J, Hensel M (2022) Data-driven design for architecture and environ-
ment integration: Convergence of data-integrated workflows for understanding and designing
environments. Spool 9(1):19–34
13. Chokhachian A, Hensel M, Perini K (eds) (2022) Informed urban environments: Data-
integrated design for human and ecology-centred perspectives. Springer, The Urban Book
Series
14. Waldheim Ch (2006) The landscape urbanism reader. Princeton Architectural Press, New York
15. Rajchman J (1998) Constructions. The MIT Press, Cambridge
16. Hensel M, Sunguroğlu Hensel D (2010) Extended thresholds I: Nomadism, settlements, and
the defiance of figure-ground. Architect Design 80(1):14–19
17. Hensel M, Turko JP (2015) Grounds and envelopes—reshaping architecture and the built
environment. Routledge, London
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1 Figure Without Ground . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Grounding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Operative Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Layered Land-Scape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Operative Layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.6 Contextual Figuration of Ground . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3 Patterns of Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1 Topological Turn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Topological Design Thinking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Patterns That Connect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.4 Organized Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4 Computing Land-Scapes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1 Performative Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2 Digital Ecology Extended . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3 Operative Extension of Nature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.4 On the Notion of Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.5 Umweltecture—Sustainable Visions Between Architecture
and Landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Epilogue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

ix
About the Authors

Pia Fricker is Professor of Computational Methodologies in Landscape Architec-


ture and Urbanism at Aalto University, Finland. She holds a doctorate degree in
Architecture and a postgraduate degree in Computer Aided Architectural Design
from ETH Zurich. Her research and teaching link urban design and landscape archi-
tecture to the field of computational design culture through the lens of emerging
technologies. Prior to her current position, she was Director of Postgraduate Studies
in Landscape Architecture at the ETH Zurich. She is a member of the editorial board
of the Journal of Digital Landscape Architecture, the Scientific Program Committee
of the DLA conference, several Peer Review Committees, and expert peer reviewer
for the International Journal of Architectural Computing, the Urban Planning Journal
and the Journal of Architecture and Urbanism. Pia Fricker has published extensively,
and her work has been exhibited, amongst others, at the Venice Biennale, the National
Design Centre Singapore, the Museum of Modern Art – EMMA, as well as at the
Helsinki Design Week.

Toni Kotnik is Professor of Design of Structures at Aalto University in Helsinki,


Finland. He studied architecture, mathematics and computational design in Germany,
Switzerland and the US and received his doctoral degree from the University of
Zurich. Before joining Aalto he taught among others at the ETH in Zurich, the
Architectural Association in London, the Institute for Experimental Architecture at
the University of Innsbruck and the Singapore University of Technology and Design.
He has been lecturing at universities worldwide as well as at museums like the
Guggenheim in Bilbao or the MOMA in New York. His practice and research work
has been published and exhibited internationally, including the Venice Biennale,
and is centered on the integration of knowledge from science and engineering into
architectural design thinking and the exploration of organizational principles and
formal methods as design driver at the intersection of art and science

xi
Chapter 1
Introduction

This book is a reflection on contemporary computational design thinking at the inter-


section of architecture, urban design, and landscape architecture, in a time marked
by complex challenges like climate change, urbanization, and population growth.
Based on a critical rethinking of the notion of ground and the relation between the
manmade and the natural environment, an understanding of architecture as regen-
erative practice is proposed where computational thinking feeds itself back into the
governing laws of nature.
In the late 1980s, the availability of personal computers and affordable software
triggered a change in architecture, landscape architecture, and related design disci-
plines, retrospectively referred to as digital turn [2]. The initiated digitalization of
the discipline has resulted in a new paradigm of computer-based design and produc-
tion often summarized under computational design. Despite some early attempts to
assess the impact on the discipline [4, 6], computational design is primarily perceived
as being of a technical nature by the majority of architects today. This develop-
ment has resulted in an ever-growing collection of computational tools for specific
design applications utilized merely for the purpose of extending and accelerating
well-established non-computational design processes [5, 7].
Such computerization of architectural production increases economic efficiency
but leaves the intellectual foundation of architecture and other design disciplines
untouched. Despite disciplinary and professional inertia, the digital turn should be
understood as part of an ongoing intellectual revolution. “Computation is influencing
research in nearly all disciplines, both in the sciences and humanities. [The computer]
is changing the way we think” [1]. This view was also stressed by Kostas Terzidis in
his early conceptual studies of computational design in architecture [8]. Contrary to
computerization, the notion of computation is not bound to the use of any machine
but existed prior to the invention of modern computers.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 1
P. Fricker and T. Kotnik, Patterns of Interaction,
SpringerBriefs in Architectural Design and Technology,
https://doi.org/10.1007/978-981-19-9083-0_1
2 1 Introduction

In the 1930s, various mathematicians, including Alonzo Church, Kurt Gödel, and
Alan Turing, started to develop precise, independent definitions of what it means to
be computable in order to find an answer to the so-called Entscheidungsproblem,
or “decision problem”, a mathematical challenge posed by David Hilbert in 1928
[3]. The concept of computation was introduced to emulate human thinking in the
process of decision-making that is as a formal language of ordering and transforming
data. The development of electronic machines, of computers, allowed for an increase
of the amount of data that could be handled and the speed of its processing. This
development gained momentum in the late 1940s and enabled the application of
computational thinking to new sets of problems. Today, the ever-increasing digital-
ization of our living environment and the pervasive availability of data allow for the
application of computational thinking to almost all aspects of life.
The resultant universality of computational thinking is affecting the way architects
are conceptualizing the world and how design interventions relate to their surround-
ings. With this, computation is changing our understanding of context on all scales
of architectural working, from the building scale up to the urban, regional, and terri-
torial scale. This observation is the starting point of the present book. It explores the
impact of computation on the design thinking in architecture not from the level of
tooling but from a conceptual perspective with focus on the fundamental notion of
context.
Taking Colin Rowe’s contextualism as a starting point, the second chapter of this
book focuses on the changing relationship between the architectural intervention
and the ground, as well as the dematerialization of this relationship through the
abstraction of digitally mediated context. Such an abstraction supports the transfer
of design methods between design disciplines and ultimately fosters the convergence
of computational design thinking between architecture, urban design, and landscape
architecture. It is argued that this convergence of computational design thinking
dissolves the difference between natural and manmade and provides the framework
for design as a contextual figuration of the ground.
The third chapter looks more closely at the process of contextual figuration through
the lens of computational design thinking. This design thinking is characterized
by the articulation of flexible relationships between entities, often denominated as
topological diagrams. It is argued that these diagrams not only govern the use of
the contextual data but are organized as a nearly decomposable web of diagrams.
From a system theory perspective, this web of diagrams is embedded in the web of
laws of nature. That means architecture is inscribed into its natural surroundings as a
manmade extension. With such an understanding, computational design intentionally
shifts away from the design of objects, of elements of consumption of environmental
resources, towards the design of interrelationships—of interactions with the envi-
ronment. The design of architecture, of cities, and of landscapes and territories is
not about the mere enhancing of environments, it is about building environments
themselves.
This reading of computational design no longer aims at sustainable design solu-
tions that minimally impact the environment but rather at regenerative design solu-
tions that actively support the natural environment and its future development. It
1 Introduction 3

aims at an architecture, an urban space as landscape, as being no less than of pros-


thetic nature. The fourth chapter presents some initial design investigations towards
regenerative architectural design that were conducted between 2018 and 2021 as part
of the design studio teaching at the Department of Architecture at Aalto University
and as joint studio teaching with Prof. Carlos Bañón from the Singapore University
of Technology and Design (SUTD) and his design studio in the Architecture and
Sustainable Design Pillar. In the subsequent epilogue, we try to relate our research
into contemporary discourse by discussing key findings with two experts in the
field: Prof. Emanuele Naboni from the University of Parma, expert in sustainable
design solutions and the digital simulation of climatic response, and Prof. Christophe
Girot, expert on large-scale landscape design and modeling methods with particular
attention to the topology of nature in and around cities.
Over the years, this approach was explored at various architectural scales ranging
from landscape design, urban neighborhoods, and city blocks, down to buildings.
Intermediate results were exhibited and discussed at a number of occasions like an
exhibition at the National Design Centre in Singapore in 2019, the Helsinki Design
Week in 2020, and the Venice Architecture Biennale in 2021. This work would
not have been possible without the support of a large number of highly committed
colleagues. We would like to thank Luka Piškorec and Kane Borg from Aalto Univer-
sity as well as Sourabh Maheshwary and Simon Rocknathan from SUTD for the
extensive computational support in studio teaching. Furthermore, special thanks go
to Tina Čerpnjak and Manuel Fonseca from Aalto University for their continuous
support in preparation of exhibitions. We also like to thank the book series editor
Thomas Schröpfer and his team from Springer for supporting our project despite
various delays. Thanks to Arley Kim for her diligent proofreading of our manuscript.
Last but not least, we would like to thank Aalto University and the Finnish National
Agency for Education for their financial support of our research and exhibitions.
Our special thanks go to our students from Aalto University and SUTD, only
through your efforts were you able to bring our ideas to life. Thanks very much
to Aino Maaike Hautala, Amirhossein Teymourtash, Anniina Norpila, Chengfan
Yang, Cheong Yi Lei, Chiuki Lai, Chong Yin Yi Christy, Choo Ee Pin, Dan Palarie,
Daniel Yong Kaijie, Egle Pilipaviciute, Eileen Wong, Elina Inkiläinen, Elizabeth
Teo Mei Qin, Faezeh Sadeghi, Fanyi Jin, Grace Teo Yu Cheng, Haipeng Wang, Han
Xianhe, Hkyet Zau Mun Aung, Ho Yu De Samuel, Hwang Jinwook, Iurii Aleksan-
drovich Shimin, Janne Jesper Keskinen, Jenna Maija Ahonen, Jiaqi Wang, Jinwook
Hwang, Joel Anthon Tiitinen, Joonas Hermanni Saarinen, Kai Antero Hakala, Kaisa
Pauliina Kiuttu, Koichi Tamura, Kyaw Zwa Thant, Lee Xuan Ying Diane, Liang
Xiuling Chloe, Loviisa Luoma, Maiju Ilona Rinne-Kanto, Maral Alaei, Marjo Linnea
Airamo, Mauricio Mari Jaelle Salas, Megan Chor Xin Yi, Mengwei Wang, Natalie
Ng Jie Lin, Nora Sønstlien, Nur Amalina Bte Md Halim, Nurul Asyiqin Zahrin,
Olga Zharkova, Petra Kaarina Suittio, Pheeraphat Ratchakitprakarn, Phoebe Kong
Li Hui, Pirita Meskanen, Quek Wen Jia Marcus, Riikka Hiltunen, Rosa Haukkovaara,
Salvador Hernandez Gazga, Sandy Low Yu Xian, Saviana Rabea Theiss, Shuaizhong
Wang, Soh Jia Ying, Solveig Sanden Døskeland, Solveig Vangen Paulsen, Sundaram
Mohan Janani, Tamura Koichi, Tina Cerpnjak, Tone Thorbjørnsen, Tong Sheng,
4 1 Introduction

Tseng Yun Ching, Tuuli Tõniste, Way Way Yun Hlwar Thon, Xin Ding, Xinyan Li,
Yang Funing, Ye Feng, Yilan Zhou, Yinan Xiao, Yiping Zhang, Yoo Fei Yi, Yuyang
Shi, Zhang Bojun, Ziyi He.
Pia Fricker and Toni Kotnik
Helsinki, December 2022

References

1. Bundy A (2007) Computational thinking is pervasive. J Sci Pract Comput 1(2):67–69


2. Carpo M (2013) The digital turn in architecture 1992–2012. Wiley, London
3. Davis M (2001) Engines of logic: mathematicians & the origin of the computer. Norton &
Company, New York
4. Lorenzo-Eiroa P, Sprecher A (2013) Architecture in formation: on the nature of information in
digital architecture. Routledge, New York
5. Menges A, Ahlquist S (2011) Computational design thinking. Routledge, London
6. Oxman R, Oxman R (2014) Theories of the digital in architecture. Routledge, London
7. Peters B, Peters T (2018) Computing the environment: digital design tools for simulation and
visualisation of sustainable architecture. Wiley, Chichester
8. Terzidis K (2006) Algorithmic architecture. Architectural Press, Oxford
Chapter 2
Convergence

2.1 Figure Without Ground

The demolition of the Pruitt–Igoe housing complex in St. Louis in 1972 is widely
recognized as marking the failure of modern architecture’s vision of urban design.
The architectural historian Charles Jenckes described this televised implosion even
as the moment “modern architecture died” [43]. The dynamiting can be viewed as an
emblematic breakthrough of the postmodern critique that had started to accelerate
and proliferate from the mid-1960s onwards [48].
One of the major influences of Postmodernism as this emerging new movement of
thoughts was Colin Rowe. His critique of the modern city and of modernistic ideas,
expressed in concepts like Ludwig Hilberseimer’s project for central Berlin or Le
Corbusier’s Plan Voisin for Paris, aimed at the “disintegration of the street and of
all highly organised public space” [53]. For Rowe, this was the result of the simple
functionalism that caused the configuration of a housing unit to be “evolved from
the inside out, from the logical needs of the individual residential unit” resulting in
the notion of a building as free-standing object.
Rowe’s reading of the modern city as a “congeries of conspicuously disparate
objects” was based on his exploration of urban morphology that he started to focus
on when he moved to Cornell University in 1962. Central to his exploration of the
urban setting and site conditions was the figure–ground diagram, a type of drawing
that illustrates in its solid–void relationship the interplay of public and private space
in an urban surrounding (Fig. 2.1). In Rowe’s urban design studio, the figure–ground
diagram became the standard format for imparting an understanding of the urban
condition [29], and the use of the drawing in urban analysis and urban design was
detailed in the seminal book Collage City, published by Rowe and Fred Koetter, a
colleague at Cornell, in 1978.
Rowe did not invent the figure–ground diagram but was one of the earliest advo-
cates of its use in urban design and planning. The diagramming technique has its
origin in the mapping of Rome by Giambattista Nolli in 1748 [63]. The so-called

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 5
P. Fricker and T. Kotnik, Patterns of Interaction,
SpringerBriefs in Architectural Design and Technology,
https://doi.org/10.1007/978-981-19-9083-0_2
6 2 Convergence

Fig. 2.1 Figure–ground diagram of Parma by Rowe and Koetter [53]

Nolli map is well known for the illustration of the city as a series of public spaces
rather than a series of objects.
Inspired by this, Rowe developed his figure–ground diagram as a solid–void repre-
sentation of built and unbuilt space that functions as a two-dimensional section of
urban space on the pedestrian level. With this type of drawing, Rowe was aiming at
opening up a dialogue between buildings and their surroundings, a way “to allow
and encourage the object to become digested in a prevalent texture or matrix of the
urban fabric” [53]. For Rowe, the figure–ground diagram became the quintessential
tool to overcome the object fixation of modern architecture and to introduce an idea
of communality as a guiding principle of urban design and planning.
… the situation to be hoped for should be recognized as one in which both buildings
and spaces exist in an equality of sustained debate. A debate in which victory consists
in each component emerging undefeated, the imagined condition is a type of solid-void
dialectic, which might allow for the joined existence of the overtly planned and the genuinely
unplanned, of the set-piece and the accident, of the public and the private, of the state and
the individual.

Rowe’s figure–ground diagram is a design tool that fosters an understanding of


the city as social space, a topic central to postmodern architecture. Furthermore,
in its gestalt theoretical grounding, the figure–ground diagram carries the seed of a
systematic problematizing of the body’s interaction with the surrounding that is of
phenomenological consideration in architecture [4]. Moreover, in the shift away from
the object towards the relationship between elements, the figure–ground diagram
paves the way towards a structuralist reading of the urban fabric [24]. Thus, the figure–
ground diagram is a design tool that relates to a range of trajectories of criticism of
2.1 Figure Without Ground 7

Modernism in architecture that have dominated the discourse for a large part of the
second half of the twentieth century.
Despite the diversity of these trajectories, they share a common concern with the
specificity of the place as well as the body in relation to this place. Both themes
were not recognized sufficiently by the Modern Movement “because of its focus on
accommodating the collective over the individual, expressed in a language of univer-
sality, both technological and abstract” [48]. Postmodernism was a counterreaction
to the perceived shortcomings of modern architecture, the abstraction and aspiration
of universality, the ignoring of local history, and the specificity of the culture of cities
as well as the individuality of its citizens.
In the case of Rowe, the postmodern critique manifested itself in the use of the
figure–ground diagram as a main tool in balancing out the modern American city
model with the historic European city model by focusing on the texture of the urban
fabric. This approach was grounded in the assumption that the urban morphology—
the texture of the fabric—is closely correlated with the sociocultural conditions in
the city [62]. Because of this, one major aspect in the exploration of historic cities
was the study of exemplary urban textures and the identification of typologies. In the
design process, the figure–ground diagram was used to analyze the urban fabric and
recognize areas of coherent urban texture and areas of rupture. Especially areas where
various textures collided were seen as necessary areas of intervention, resulting in
the proposal of ambiguous, hybrid configurations often based on the collage-like
merging of ideal urban typologies in order to integrate into the neighboring urban
textures.
The figure–ground diagram was thus used in extracting urban textures that promul-
gated patterns of regularity and weave them together within the design process. It is
in this sense that Rowe’s approach can be seen as being contextual because the term
context has its etymological roots in the Latin contextus meaning “joining together”,
which originally was the past participle of contexere, “to weave together” [21]. Rowe
established Contextualism at Cornell University as an urban design approach that
attempted to overcome the object of the Modern Movement by intertwining it with
its neighborhood on a morphological level.
The object itself, however, remained untouched. In other words, Rowe did not
question the relevance of modern concepts in the design of buildings but rather the
validity of the concepts on the larger scale of the urban [62]. In this, his critique of
modern architecture differed very much from that of Robert Venturi, architectural
theorist, and central figure in the Postmodern Movement in architecture. But in
essence, Rowe’s urban design approach clearly was in line with Venturi’s manifesto-
like description of the postmodern critique that he formulated in his publication
Complexity and Contradiction in Architecture [58]:
I speak of a complex and contradictory architecture based on the richness and ambiguity
of modern experience, including that experience which is inherent in art... I welcome the
problems and exploit the uncertainties... I like elements, which are hybrid rather than “pure”,
compromising rather than “clean”, ...accommodating rather than excluding... I am for messy
8 2 Convergence

vitality over obvious unity... I prefer “both-and” to “either-or”, black and white, and some-
times grey, to black or white... An architecture of complexity and contradiction must embody
the difficult unity of inclusion rather than the easy unity of exclusion.

Rowe’s Contextualism was not the only architectural position that emerged during
Postmodernism and tried to develop “the difficult unity of inclusion”. Most well
known is the existential reading of place introduced by architectural theorist Chris-
tian Norberg-Schulz in his study Genius Loci: Towards a Phenomenology of Archi-
tecture [49]. But what is common to all these postmodern contextual approaches is
an emphasis of the social and cultural dimension of architecture as a main driver of
the design process.
The environmental dimension, or the physical context and its impact, never played
a significant role in the design thinking of postmodern architecture. The historic
function of buildings as environmental modulator was progressively relegated to
a secondary place in the discourse and in practice effectively handed over to the
emerging profession of mechanical and electrical consultants [31]. Aspects like sun
shading, improved air flow, or the collection of rain water had been relevant param-
eters in the design of the modern city [37], like in Hilberseimer’s principles of urban
planning presented in The New City [34] or in Le Corbusier’s idea of a vertical garden
city visible in the Ville Radieuse of the Unité d’Habitation in Marseille. Based on
this, the physical environment was perceived as being part of the domain of a scien-
tific rationality, of a technological operativeness that the postmodern critique tried
to overcome [30].
Such a disregard of environmental thinking is also apparent in the figure–ground
diagram. The drawing reduces the city to the building mass and the space between the
buildings. In some cases, elements of the infrastructure are added as line drawings
in the diagram. In Rowe’s design approach, the view onto the city is reduced to the
manmade; information on the natural, physical environment is missing. Without even
the smallest topographic hint, the city seems to float above a flat ground that functions
as a neutral background: an environmental tabula rasa. The figure is without ground!
A comparable conceptual neutralization of the ground, or detachment from the
territory, is inherent to the Modern Movement as well. The literal emancipation of
the architecture from the ground, like in the Villa Savoye by Le Corbusier or the
Farnsworth House by Mies van der Rohe, not only creates a new, elevated level zero.
It also underlines the independency of the building from the given conditions of the
ground; the universality of the architecture; and, with this, its quality as an object.
“Through this physical, programmatic, and semantic emptying of the ground the
context mutates ultimately into that mass without qualities, which, in the form of the
tabula rasa, was to become the prima materia of modern urban planning” [54].
The figure–ground diagram, thus, is the key tool in Rowe’s critique of Modernism
and, at the same time, its locus of continuity. Such absence of a rethinking of the
ground is not limited to Rowe’s discourse only but exemplary for other contextual
explorations. New definitions of context were introduced into architecture in the early
1950s and started as studies of the perceptual form of the city, most prominent in
Kevin Lynch’s The Image of the City [41], but soon departed towards a more inclusive
2.2 Grounding 9

definition that encompassed cultural context and social practice within a postmodern
discourse [13]. Postmodern variations of the notion of context, like Norberg-Schulz’
genius loci or the socioplastics of the Smithsons’, the idea of ambiente of Rogers as
well as Rossi’s locus, all share Rowe’s ambition to dissolve the modernistic object
by including existing cultural and social circumstances. With this, postmodernism
provoked an epistemological shift of the figure resulting in a centering on the human
perspective by means of conceptualizing the social, cultural, and political relation-
ships between people and their surroundings. Within architecture, this may have
caused a rethinking of the figure, but it did not cause a rethinking of the relation
between the figure and the ground.
The ground is the blind spot of postmodern architecture! This is a blindness it
shares with Modernism.

2.2 Grounding

Modern Brazilian architecture illustrates that this conceptual neutralization of the


ground is not inherent to the scientific rationality or technological operativeness
of modernism but rather a constitutive part of the main narrative of the Modern
Movement and its postmodern critique. Rooted in Le Corbusier’s ideas, modern
Brazilian architecture is characterized by the reconciliation of a search for a unique
identity along with a concerted effort of the country to engage in contemporary
debates [33]. The experience of the tropic, thereby, played a central role in this
search for identity with modernity as a way to rethink nature and landscape with a
new rationality [46]. This is particularly visible in the work of Paulo Mendes da Rocha
who believes in architecture as a means of constructing a new landscape as “man’s
ability to transform the place where he lives”. His design for the Brazilian pavilion
for the Expo 1970 in Osaka is a quintessential articulation of this idea (Fig. 2.2).
The pavilion is reduced to a flat roof that seems to float above an undulating
landscape. The large concrete canopy of around 30 m × 50 m rests on three columns
and a gate-like arch with the columns countersunken into artificial hills. With sunlight
filtering through the gridded roof structure, the canopy floats like a light surface that
touches the landscape very gently. The architecture and the landscape form a unity
in dialogue, a new place for life to unfold. This interaction between the engineered
structure and the ground shows a certain poetic characteristic for the work of Mendes
da Rocha that transcends mere technical adequacy. Creativity and engineering are

Fig. 2.2 Paulo Mendes da Rocha, Brazilian Pavilion, Expo 1970, Osaka, Japan, sketch, 1969 [50]
10 2 Convergence

used to occupy a territory in a way that tests the boundaries between natural and
artificial, and in essence construct a new landscape [46].
Together with João Vilanova Artigas, Paulo Mendes da Rocha was the main
protagonist of what is commonly known as the Paulista School of Brazilian Archi-
tecture. From 1957 onwards until the mid-1970s, Vilanova Artigas and Mendes da
Rocha contributed a number of important works of outstanding quality that resulted
in the spreading and adaptation of their ideas throughout the country [33]. But despite
his importance for the development and recognition of architecture in Brazil, Mendes
da Rocha did not perceive himself as a Brazilian but rather as a modern architect:
“Being an architect is not just about where you are. Architecture is universal. …
Being an architect is a matter of knowledge—you explore the place and interpret
how to respond to a particular site and situation. Water is water, gravity is gravity,
and sunlight is sunlight. It is the same everywhere. … Architecture that is done here
can only be interesting when it possesses a universal dimension. There is no such
thing as Brazilian architect” [6].
This belief in the universality of architecture and the engineered rationality of
the design connects the work of Mendes da Rocha with the Modern Movement.
But the rationality of the design is not developed out of functional considerations.
For Mendes da Rocha “architecture does not desire to be functional; it wants to be
opportune” [6]. Universality, rationality, and technological operativeness in the work
of Mendes da Rocha are an expression of modernity as an attitude: an attitude towards
the world and one’s own being-in-the-world. This attitude is equated with individual
freedom and its instrumental aim of changing and adapting the world. With Mendes
da Rocha, this changing and adapting of the world becomes visible in the entangling
of architecture and site into a new constructed landscape.
This contrasts starkly with the conceptual neutralization of the ground in the
Modern Movement. In Mendes da Rocha, we see a figuration of the ground as an
essential part of the design, a move from the dichotomy of the figure–ground towards
a liaised figure–figure relationship between the building and the ground. With his
work, Mendes da Rocha points towards a reconceptualization of the ground within an
architectural discourse that started to emerge as a counterreaction to postmodernism
in the late 1980s. According to Peter Eisenman [18].
No other discourse occupies ground in the way architecture does. … there is a literal ground,
there is also an ideal ground, a conceptual ground, an abstract ground if you want, that deals
with what architecture is about today and in the future, that is the ground of the urban, the
ground of the rural, the ground of any institution. As we build on this Earth, how, in what
way, ground is conceptualized becomes important.

Modernism and postmodernism were focused primarily on abstract and ideal


grounds, on cultural and social grounds, respectively. In the conceptualization of
architecture, the literal ground was neglected or simply taken for granted. This
changed with the winning competition entry for the design of Parc de la Villette
in Paris by Bernard Tschumi in 1982.
The competition brief by the French government asked for an urban park for the
twenty-first century. Tschumi’s answer to this request was a transformation of the
2.2 Grounding 11

prevailing idea of a public park as an escape from the city into a continuation of the
city and place of social interaction. He achieved this by abstracting the landscape into
an open surface overlaid by a regular grid of so-called follies, red structures without
function that provide orientation to the user of the park. Additional objects in the park
are curving paths and landscaped green surface patches as well as cultural institutions
like museums and concert halls that attract people and promote interaction. The
design of Tschumi is an architectural interpretation of landscape and redefines the
urban park as an infrastructural element of the city.
Parc de la Villette not only activated a rethinking of the relation between landscape
and the city and the emergence of “landscape urbanism” as a new discourse within
architecture promoted by Corner [11], Mostafavi [47], Waldheim [59], and others.
It also started a rethinking of the urban context by means of landscape strategies in
architecture.
Bernard Tschumi was instrumental in this rethinking of the urban context. During
his deanship at Columbia University between 1990 and 1995, he assembled a group
of young and ambitious architects, including Greg Lynn, Jesse Reiser, and Stan Allen,
and encouraged them to use the design studio as a laboratory. Around the same time,
computers were introduced into architecture at a larger scale and started to open
up new possibilities in design. Within this atmosphere of friendly competition and
debate, new concepts arose “at once from an intuition about urban context—the city
as a dynamic field, mapped and understood on site—but also from an intuition about
harnessing the power of the computer as an abstract calculating machine” [2].
One of these new concepts central to the rethinking of the relation between archi-
tecture and the urban context is the so-called “field condition”, formulated by Stan
Allen in his article From Object to Field [1]. Based on the non-figural response to
context in examples like Le Corbusier’s Venice Hospital and Rafael Moneo’s analysis
of the Mosque at Cordoba or the reading of the American city as a pragmatic accu-
mulation of small variations that “unpack the ideality of the grid” [1], the concept
of field emerges as a production of difference out of local “part-to-part relationship
while maintaining an overall coherence” [2].
For Allen, a field condition is fundamentally an urban concept that emerges
“out of the desire to pay close attention to the intricacy of the contemporary city,
and to respect the capacity of the city—and its citizens—to produce complexity
and difference on their own terms” [2]. Such an understanding of the city as a
self-organizing entity governed by local interconnectivity carries in itself already
a biological connotation.
Despite this rather obvious conflation, the concept of field condition was not
associated with ideas of landscape from the beginning. But over time, this rela-
tion surfaced and resulted in a number of publications on landscape strategies in
architecture published between 2001 and 2011 with programmatic titles like Land-
scrapers [8], Topographical Stories [39], Groundscapes [54], Landform Building
[3], or Groundwork [5]. These books illustrate the increasing awareness of the rela-
tionship between landscape and architecture in a series of built projects with the
aim of repositioning the conventional understanding of architecture and landscape
12 2 Convergence

within the domain of contemporary ecological theories without giving any conclusive
arguments [35].
From this point of view, these books have to be seen as a motivation for future
research into the relation of architecture and landscape—of figure and ground—rather
than conclusive summaries of a development.

2.3 Operative Surfaces

A starting point of such research is a more thorough reading of Stan Allen’s field
condition which has originated from an exploration of the development of the Amer-
ican city in contrast to the management of territory by the Jeffersonian grid [1].
Following Colin Rowe, Allen observes that the grid is “convincing as a fact rather
than an idea”. American cities of the Midwest and the West are characterized by
the accumulation of small variations to the Jeffersonian grid, perturbations that are
smoothly accommodated within the overall order. What can be observed in the Amer-
ican cities is a “close attention to the production of difference at the local scale, even
while maintaining a relative indifference to the form of the whole. … Variation and
repetition—individual and collective—are held in delicate balance” [1].
In contrast to the Jeffersonian grid, Allen’s reading of the grid is not a top-down
implementation but rather a bottom-up process of local adaptation with each cell of
the grid attaching to the boundaries of the immediate neighborhood and, at the same
time, adjusting itself to the local conditions on site. It is this mechanism of local
interaction that Allen emphasizes in his study of Le Corbusier’s Venice Hospital:
“The placement of blocks establishes connections and pathways from ward to ward,
while the displacement of the blocks opens up voids within the horizontal field of
the hospital. There is no single focus, no unifying geometric schema. … The overall
form is an elaboration of conditions established locally” [1].
Viewed as a field condition, the individual cell of a grid or individual block of
the hospital becomes the individual element within a set of elements not placed
anymore onto a neutral ground but finds its relative position through an interaction
with its local surroundings, that is, with neighboring elements and the ground. A field
condition therefore describes a triggering effect of organization on a whole set of
elements, the horizontal phenomena of figuration of the whole out of the interaction
of smaller individual parts similar to schools of fish, swarms, or crowds.
As a consequence, a field condition is a surface-oriented design strategy with the
surface defined as collection of elements that interact with one another. This local
interaction is informed by local ground conditions as well. This means the ground
has to be seen as an active surface of information interacting with the surface of
elements. Stan Allen’s field condition, thus, transforms Colin Rowe’s static figure–
ground relation into an active interplay of two surfaces, the figure–surface of internal
interaction of elements and the ground–surface of external information acting upon
the figure–surface.
2.3 Operative Surfaces 13

Fig. 2.3 SANAA, Rolex Learning Center, EPFL, Lausanne, Switzerland, 2010 (Image: Alain
Herzog/EPFL)

The Rolex Learning Center at the École Polytechnique Fédérale de Lausanne


(EPFL), designed by SANAA, exemplifies such a transformation of the figure–
ground relation into an interaction of two surfaces (Fig. 2.3). The Learning Center,
a programmatic mix of library, bistro, conference center, meeting and exhibition
spaces, and work places for students and scientists, is located at the edge of the
EPFL campus, a flat site with views to Lake Geneva and the surrounding mountains.
The building is integrated into the organizational grid-pattern of the university and
covers an area of around 166 m × 122 m. One of the main purposes of the Learning
Center is to connect the EPFL campus with the neighborhood and with it integrate
the campus better into the surrounding neighborhood.
In SANAA’s design, this motivates the lifting up of the building from the ground
in order to enable passages through the site without the need to enter the Learning
Center. In addition, specific functions are intentionally elevated for better framing
of views of the surrounding mountains and lake [35]. Circulation patterns and view-
points of the Alps are external information of the ground that is acting on the figure–
surface as an imaginary force causing topographic deformation. At the same time,
questions of walkability and usability inform the curvature of the surface and define
local rules of interaction for the surface patches of the figure–surface. These two sets
of rules result in a vibrating, undulating figure–surface into which elliptical holes are
punched that support the functional distribution within the surface, bring light to the
ground level below the building and enabling both visual and physical connection
between the inside and the outside.
14 2 Convergence

The Rolex Learning Center at the EPFL illustrates how the figuration of the
figure–surface is the result of the interaction of a number of formal operations acting
upon each of the surface patches of it. This operativeness is implied by the given
field condition, it is a creative translation of the conditions of the site into geometric
transformation. Field conditions, thus, establish a performative design approach as
an abstract machine that Deleuze and Guattari describe as diagrammatic; in essence,
a thinking that “operates by matter not by substance; by function, not by form” [15].
Computational means are required to activate this diagrammatic abstract machine
of the field condition and achieve control over the multitude of interactions inherent
to the design intention. In retrospect, it is not a coincidence that field conditions
appeared at the same time when computers were introduced into the design studio
[2]. The digitalization of the discipline, the technological advancement of design
and fabrication through computation only made it possible to construct the hilly
landscape of the figure–surface of the Rolex Learning Center [9].
From an architectural point of view, the deformation of the figure–surface into a
hilly landscape introduces differentiation and with it a spatial dialogue between the
two surfaces of the figure and the ground. The differentiation results in a subdivision
of the ground into a sequence of smaller spaces that sometimes open up to the sky,
sometimes have a grotto-like quality, sometimes are transitional spaces, and some-
times are enclosed courtyard spaces. The field condition, thus, not only introduces
an abstract, operative dialogue between figure–surface and ground–surface, but in
SANAA’s design, this abstract interaction is translated into a spatial one: a dynamic
in between that overcomes the neutrality of the ground despite its untouched flatness.
A dynamic in between that spatially resembles the dialogue between architecture
and nature, between figure and ground, in Mendes da Rocha’s Brazilian Pavilion
for the Osaka Expo. In the pavilion, the flat roof functions as reference surface
that guides the physical manipulation of the ground. The countersunk columns act as
points of attraction along which the flat ground is transformed into a rolling landscape
that shapes the plaza underneath the roof and guides the flow of people within the
site. These operations of material accumulation introduce a ground differentiated by
slope, levels, widening, and closing of gaps and other geomorphic phenomena that
can be activated for architectural purposes.
In the Brazilian Pavilion as well as in the Rolex Learning Center, we can see a
spatial dialogue between figure–surface and ground–surface caused by operations of
pulling and pushing acting upon one of the surfaces. In both cases, the flatness of
the second surface acts as stable reference plane and the duality of the relationship
between figure and ground is not questioned.
This is not the case anymore in the Yokohama Port Terminal by Foreign Office
Architects (Fig. 2.4). The terminal is a 430-m-long infrastructure project that
combines passenger traffic, car traffic, and a public garden into a complex flow of
spaces where the relationship between interior and exterior and the flow of passen-
gers and the flow of people in the gardens is constantly questioned. This happens by
the simultaneous figuration of the figure–surface and the ground–surface resulting
in an undulating terrain of connecting ramps and terraces.
2.3 Operative Surfaces 15

Fig. 2.4 Foreign office architects, port terminal, Yokohama, Japan, 2002 (Image: Satoru Mishima)

Operations of folding and slicing are no longer limited to one of the surfaces but
are acting on both at the same time, influenced by the flow of people between ferries
and the city. The operations become an expression of the agency of infrastructure.
Through the use of the same materialization, the distinction between the different
programmatic functions is further blurred. In such a design, the notion of figure and
ground is obsolete. Rather there are a number of surfaces that interact with each
other by means of geometric operations that results in a multilayered constructed
landscape.
It is only consequential that the building rests on water and does not provide a
traditional reading of figure and ground. According to Stan Allen, the Yokohama Port
Terminal “is perhaps the most convincing realization of an architecture invested in the
idea of landscape techniques working at the scale of a building. Indeed, Yokohama
is nothing if not a constructed landscape. … The architects have literally constructed
a new site” [3].
This new site is the result of dissolving any hierarchical opposition between figure
and ground. Instead, figure and ground are seen as two layers of information that
evoke operations of interaction with the aim of generating architectural space. The
layers of information define the formalized context out of which the spatial config-
uration emerges by means of well-defined correlations and transformations. With
this, architecture has to be seen as part of its own context, as part of the ground it is
placed, as a constructed land—as landscape.
A closer examination of Stan Allen’s field condition seems to confirm the associ-
ation of the concept with the idea of landscape visible in the abovementioned series
of publications on landscape strategies in architecture titled Landscrapers [8], Topo-
graphical Stories [39], Groundscapes [54], Landform Building [3], or Groundwork
16 2 Convergence

Fig. 2.5 Field condition as operative entanglement of figure–surface and ground–surface and three
basic modes of interaction: a ground-active, b figure-active, c interactive

[5]. What seems obvious in all these titles, however, is an understanding of land-
scape as the other, as something that is outside of architecture but consumed by it.
In essence, it is a perpetuation of the figure–ground relation.
What the concept of field condition really evokes is not a convergence between the
architecture and landscape as a discipline but rather a convergence of architectural
design and design methodology in landscape architecture: a convergence of design
thinking based on technology [17]. This convergence is characterized by the operative
interplay of layers of information that enables a computational process of informed
landscape design (Fig. 2.5).

2.4 Layered Land-Scape

Such an operative interplay of layers of information is even more evident in the


development of landscape architecture. In 1969, the influential book Design with
Nature by Ian Lennox McHarg laid the foundation for a design and planning theory
applied within landscape architecture based on interactive, nature-based processes
as the elementary starting point of the design and planning process. According to
McHarg, “the basis of the method is constant for all case studies—that nature can be
considered as interacting process, responsive to laws, constituting a value system,
offering intrinsic opportunities and limitations to human uses. Now better armed, we
can take our knowledge of nature as process and apply this to a problem to discern
the place of nature in a metropolitan region” [44].
Based on processes in nature, local conditions are understood as an element of
global scale and therefore stand in direct correlation with system thinking princi-
ples discussed in the realm of cybernetics. Similar to the triggering effect of field
conditions (see Sect. 2.3), McHarg’s approach implies the horizontal phenomena of
figuration of the whole out of local interaction of parts.
McHarg defined a new interdisciplinary strategy of understanding nature as a
dynamic process, one that strongly impacted the field of education and practice, as
well as the theoretical discourse in relation to ecological issues and the development
of urban and landscape ecology. For McHarg, site design is not simply a matter of
2.4 Layered Land-Scape 17

spatial shape, form, and experiential impression, but more the adaptation of various
ecological forces in time to shape a more dynamic and complete whole [12]. This
holistic perspective allowed for the development of new planning strategies based
on an understanding of a place and its unique parameters, such as land morphology,
soils, stream patterns, plant association, wildlife habitats, and land use. Based on this
interdisciplinary perspective, McHarg developed the so-called Layer Model Strategy
(Fig. 2.6), a precursor of the Geographical Information Systems (GIS) [32].
While the figure–ground method established in architecture defined the relation-
ship between built and unbuilt spaces for city planning, McHarg focused on the
complex dynamic vertical connections of layers of the physical ground by advocating
the use of the overlay method as an objective instrument. In McHarg’s Layer Model,
each layer is perceived as dynamic set of information. By simulating nature-based
processes, the evolution of information within layers could be extended into the
future. This time-based approach incorporates additional layers of future “system

Fig. 2.6 The Layer Model by Ian McHarg visualizes the interaction between the mapped site-
specific conditions and phenomena and its interconnections through clustering and incorporation
of time [44]
18 2 Convergence

components”, like plants, animals, and humans, all dependent on the overarching
influence of micro-, mezzo-, and macroclimatic conditions.
Design with Nature discusses a very timely strategy of living with the forces and
flows of nature, in opposition to the top-down planning mechanism of the time. The
scientific research into the underlying complex rules of natural processes provides
a knowledge base for designing within a planetary network. Although this evalu-
ation process took place in an analog manner, it was revolutionary as it addressed
the complexity of different systems previously analyzed in isolation. Even though
McHarg understood the environment as a dynamic process, his chosen mode of repre-
sentation can be critically viewed as plotting and ranking “of natural phenomena on
static maps” [32] without focusing on iterative strategies.
This shortcoming in the interaction of layers has been reinforced by the devel-
opment of computational tools. Inspired by the methods of Ian McHarg and due to
increased pressure to find answers for current urban challenges, the Harvard Labora-
tory for Computer Graphics (LCG), founded in 1966, focused increasingly on the area
of workflows for spatial analysis and visualization [20]. The main objective was to
design a computer-mapping program called Synagraphic Mapping System (SYMAP)
(Fig. 2.7). This project was initiated by the leadership team of Harvard University
to convert traditional, analog mapping into a comprehensive tool, but already from
1967 onwards moved onto environmental topics and integrating ecological knowl-
edge [60]. The project on the regional development and conservation of the Delmarva
Peninsula by Carl Steinitz in 1967 was the first application of GIS in making a design
for a large geographical region. The project prepared computer programs in Fortran
and used SYMAP to produce the first coded evaluation models for the future land
use [56]. In the project, complex site-specific data was visualized graphically and
allowed for the development of a first computationally driven data-informed design
strategy.
In 1969, Jack Dangermond, a former member of the Harvard Lab, and his
wife Laura founded the Environmental Systems Research Institute, Inc. (Esri). The
consulting firm applied computer mapping and spatial analysis to help land use plan-
ners and land resource managers make informed decisions. The company’s early

Fig. 2.7 Overview of basic map types generated by SYMAP. The main functionality of the program
was to translate site-specific information into vector content that was then stored as data [55]
2.4 Layered Land-Scape 19

work demonstrated the value of GIS for problem-solving “by a clear vision that a
mapping and analysis framework could provide a deeper understanding of our world
and help us design a better future” [14]. Esri went on to develop many of the GIS
mapping and spatial analysis methods now in use, focusing on spatial and landscape
planning within a specialized software environment disconnected from the design
environment. The integration of geospatial data packages for simulation and analytic
use in the design process, however, was only integrated into commonly used design
software over the past few years.
Esri’s market domination in the use of GIS data can be seen in direct correlation
with the stagnancy of experimental use of data in landscape architecture. More-
over, the use of data within design using the established Layer Method Strategy
was reduced to pure analysis and simulation use of existent data, but did not take
into account a dynamic interaction with the design process [61]. Due to the black-
box character of specialized GIS tools, designers were not focusing anymore on the
scientific process-oriented component, but rather on the outcome of the simulation.
As a consequence, the understanding of complex interactions through layering
and mapping of local information, combined with design intention, has been one of
the most influential design methods in the field of landscape architecture to date. The
Fresh Kills Park on Staten Island, designed by Field Operations between 2001 and
2006 (Fig. 2.8), demonstrates the similarities in the developed methodology to the
approach of McHarg’s Design with Nature.

Fig. 2.8 The proposed design scheme of the Lifescape project builds upon the Layer Cake Strategy
and “separates between historic processes of waste and wetlands and the proposed new program,
habitats, and circulation above” [19, 196] (Image: New York City Department of Planning and Field
Operations)
20 2 Convergence

The thoughtful creation of these layered categories of information is the result of


deep studies on the macro- and micro-scale of the environment and its ecological
processes and has been celebrated in the project through a vertical separation between
the challenging conditions of the former landfill site and the newly planned park [51].
The history of the site, however, has been fully separated from the future development
through the added functional separation layer of “soil cover”. Even though Field
Operations focused on a new reading of nature, understood as an inhabited field and
not anymore as an object, this layered design approach has been critically viewed.
The disconnected formal reading of the components is missing a methodological
approach to reconnect and combine data, bringing different relationships and sub-
systems into an ecological discourse in order to evaluate and represent the conditions
[19].
With its independent layering, the design of Fresh Kill Park resembles Tschumi’s
design strategy for Parc de la Villette (see Sect. 2.2) and his architectural interpre-
tation of landscape in the Field Operations is an attempt to allow information from
different clusters to interact with one another in the hope of activities to emerge out
of the superimposed organization. But as Michael Dring and Ed Wall have pointed
out, “instead of vertically layering up programs and processes onto the site, existing
processes can be understood as redirected, adapted, and combined. The continua-
tion of McHarg’s analytical layering approach into a methodology through which
new design solutions are formed offers clarity in a process of designing landscapes.
However, this conceptualization of a proposed landscape separated into layers can
leave it insufficiently resolved and is “reinforcing static hierarchies of unconnected
processes” [19].

2.5 Operative Layers

The greatest design potential lies exactly in the interplay of the various layers of
information. This is demonstrated by the Botanical Garden in Barcelona by Carlos
Ferrater (Fig. 2.9), constructed between 1989 and 1999 on a former waste landfill,
where the exchange of layered information motivates a design method of modeling
a new topography and with it an “architectonic designing of the landscape” [22].
The design concept of the Botanical Garden was rooted in a deep understanding
of the local site-specific conditions and guided by the “morphological, topographical
and topological conditions” [22]. The layout of the garden considered botanical and
ecosystem questions related to varieties of flora from a range of Mediterranean areas
synthesized into a triangular grid over the terrain determined by the three main
directions of the contour lines. Other subdivisions of the grid took into account
notions of planting, drainage, and orientation towards the sun but also accessibility
and walkability. All this information cumulated in concave or convex pairs of walls
that frame the triangle of the grid and fracture the constructed landscape.
The developed methods focus not on developing a representation technique for
dynamic data or a layered representation of functions and local conditions, but
2.5 Operative Layers 21

Fig. 2.9 Barcelona Botanical Garden by Carlos Ferrater, Bet Figueras and José Luis Canosa [23]

rather on a sensible way of displaying site-specific perception in relation to spatial


conditions and future system interactions.
In contrast to the layered juxtaposition of information in the design of Fresh
Kills Park, Ferrater’s design of the Botanical Garden introduces a new surface, the
triangulated grid, as a target layer. Similar to the shaping of the figure–surface of
the Rolex Learning Center, the target layer reacts to layered information through
deformation (compare Sect. 2.3). Conceptually, the design describes a figuration of
the figure, a figure-active entanglement of the triangulated grid as figure with the
layered ground information (Fig. 2.5b).
The site is dealt with as a vector field, allowing for a rule-based manipulation of
the target surface. The ground acts as a continuous field, which allows for geometrical
manipulation according to the envisioned future processes. The careful designed path
system follows the basic system of the vector field and relates to the morphological
and topographical conditions, creating zones for the “harmonious distribution units
of vegetation” [23] through geometrical operations like the clustering of vertical
movement of the connection points, according to the needs of the vegetation. This
system approach is connected to a deep understanding of ecological principles and
theories of networks. Environmental conditions like sun direction, flow direction of
runoff water as well as the needs of the future users and actors of the park are under-
stood as system components and form the basic principles for the transformation of
the grid.
From a structural perspective, the arrangement of the two distinguished actors
of the design, the structuring and passive “network of paths, walls and intersti-
tial spaces”, and the active and process-oriented “vegetation and plant community”
allow for a dynamic interaction and transition, offering the “natural order of the new
22 2 Convergence

landscape” [23]. Designing with the principles of nature automatically showcases


a process-oriented approach, as the shaping of the terrain and the arrangement of
the vegetation are designed within one system component, relating to site-specific
environmental conditions, as well as to the needs of the users of the park. These two
actors of the design are interwoven by means of “geometry as a way of approaching
landscape and urban form” [22]. The figuration of the triangular grid constructs
an intrinsic order that can help to “recognize the place and pick out its hidden
needs, forcing its authentic condition to surface” [22]. A similar instrumentaliza-
tion of geometry for the visualization of the intrinsic order is also the main driver
of the renaturation project of the river Aire in Geneva by SUPERPOSITIONS, with
Georges Descombes and ADR (Fig. 2.10). The renaturation project aims for a new,
immediate understanding of natural processes, creating a linear “rivergarden”, which
reacts to ecological and riverine processes in exchange with recreation and pedestrian
circulation, acting as a strategic component for further urban development [16, 57].
Originally, the Canton Geneva aimed for a restoration of the Aire River from its
canalized banks, which is situated in an area fragmented by suburbanization [52].
The design team took the challenge of working with the dynamic forces of the water,
proposing to keep the original canal—understood by Descombes as a “reference
line”—and adding a parallel second riverbed that acts as a floodplain [16]. The
design’s dialogue with the engineered history of the river symbolizes an act of control,
thereby opening up the site to “deeper questions about nature, temporality, and the
making of landscape” [52].

Fig. 2.10 Renaturation Project of the river Aire (2002–2016): transformation stages of artificial
pattern interaction with the dynamic forces of water (Image: Superpositions)
2.5 Operative Layers 23

Contrary to the Botanical Garden, the renaturation project does not aim at a static
form for the project but rather at an illustration of processes in nature. With this, the
project adds to the discussion on “process versus product” initiated by Marc Treib
[57] in an attempt to extend the discourse on “forms follows function”, still prevalent
in the design of the Botanical Garden. It also relates to a discussion on landscape as
a design driver following Girot [28]:
Landscape is evidence of an accrued intelligence of place through topological transformation
and an exchange of techniques, beliefs and actions. The location of a particular terrain,
its specific arrangement within borders and the manner in which permanent and seasonal
vegetation is handled touch not only on the constructed materiality and pragmatism of an
individual culture, but also on the aesthetic and symbolic level as a source that nourishes
beauty and reverence for nature.

In the river Aire project, the geometry inherent in the flow of water is made
visible through the successive transformation of the gridded flood plan. Similar to
the roof of the Brazilian pavilion of Mendes da Rocha (see Sect. 2.2), the grid
functions as neutral, static reference plan against which the movement of the river
is made visible. The resulting figuration of the ground (Fig. 2.5a) celebrates the
capability to aesthetically demonstrate morphological processes, guided by a lozenge
pattern of small depression [36] and described by Decombes as “fluid morphology”
[57]. The initial abstract pattern acts as new topography, allowing for an interaction
between forces and processes, including, as stated by Anita Berrizbeitia, “history as
an unfolding, multilayered process” [7].
The inherent logic of the river dominates the design process and ultimately stages
the river itself as a “co-designer”, addressing fundamental questions of transforming
landscapes and the understanding of nature [10]. For Treib, “the new river is not a
natural product, it is a human construct that has interwoven intellect and talent with
natural forces and materials. These factors have conspired in creating an instructive
and experientially rich landscape in a state of constant evolution—but an evolution
within a framework established through design” [57].
Such an understanding of landscape as a human construct in a state of constant
evolution has been taken to the extreme by Atelier Girot in the design of the Sigirino
Mound (Fig. 2.11), a design exemplified by an in situ computational design process
influenced by ecology and supported by new topological methods [28]. According
to Girot, “the goal was to make a landscape that could fit within a highly sensitive
cultural setting and a vulnerable ecological environment…. To help in the shaping
of an optimal form, the project team applied advanced modeling and visualization
techniques based on point clouds. This enabled the designers to test and present a
variety of ecological options during the approval process” [28].
The landfill design project fully embraces an iterative process-oriented approach,
taking the local parameters as conditions for a performative design concept based
upon a complex interplay of layers. The design challenge was to amass around
3.7 million cubic meters of inert stone material from the Alp Transit high-speed
railway tunnel in Switzerland and integrate it into the surrounding landscape [25].
The material was to be applied on site to form the Sigirino Mound, one of the largest
artificial mounds in Switzerland [26].
24 2 Convergence

Fig. 2.11 Sigirino Mound under construction. Computational design process: the applied point
cloud technology builds the digital backbone of the developed iterative design process (Image:
Atelier Girot)

The main goal of the design was a respectful interaction with the surrounding
environment, thus capable of demonstrating its own characteristics. The in situ
design process was based upon real-time data capturing, visualized through point
cloud technology, which allowed for an in depth understanding of the geological and
morphological behavior of the applied material. The mechanical stabilization of the
excavation material was one of the main challenges to be solved through rethinking
commonly used design and construction methods in “order to facilitate plant growth
on inorganic substrata” [61].
The structural property of the material is monitored and translated into a design
language of material behavior in relation to ecological processes like the inhabitation
by vegetation. The performance-based approach of the design works hand-in-hand
with the properties of the excavated material and envisioned future landscape. The
overall geometrical logic of the artificial mound, the width of the path-like serpentines
as well as the stepped and faceted morphology of the layers are informed by the
construction methods.
Even though this project could be understood as a layer-based design approach
from a visual point of view, the methodology goes beyond the traditional reading of
2.6 Contextual Figuration of Ground 25

surface-based properties, demonstrating a new topological design language, melding


ecology, design, and computational design thinking. The constant flow of informa-
tion between the layers during the construction process and the impact of these
information on the allocation of material describes a design method that resembles
the interactive interaction of figure and ground in the design of the Yokohama Port
Terminal (see Sect. 2.4). Like in the Yokohama project, the notion of figure and
ground does not describe the process sufficiently. Rather, there are a number of
surface layers that interact with each other by means of geometric operations that
result in a multilayered constructed landscape (Fig. 2.5c).
The iterative exchange between geo-referenced point cloud models, visualized in
highly accurate scale models, supported the iterative design process and allowed for
performance analysis in relationship to drainage, planting, and visual impact. This
feedback loop between local data interaction and overall design intention related to
the concept of landscape topology, which “does not base its question on the finished
‘image’ of a landscape’s possible power of synthesis; it analyses the process behind
the power of reflection itself” [27].

2.6 Contextual Figuration of Ground

The Sigirino Mound is an extreme example that illustrates the general intention
of architecture and landscape architecture of transforming given conditions, locally
creating new environments. Both architecture and landscape architecture are material
practices that attain social, cultural, and ecological relevance through the articulation
of material arrangements and structures [45] with the articulation based on two groups
of operations: the rearrangement of existing material on site and the adding of new
material to it.
It is the application of these two groups of operations that results in a figuration
of the ground. The ground is no longer understood as a simple surface, as the neutral
background to the architectural figure as in Rowe’s figure–ground diagram, but rather
as a pre-existing material arrangement, an amalgamation of natural and manmade
efficiencies that get transformed by a sequence of operations. Such a notion of ground
is a procedural one that does not differentiate between natural and manmade envi-
ronments but unites them into the unique environment that forms the backdrop—the
ground—for our daily life on earth.
Such a unifying understanding of the ground is fostered by the operative nature
of a computationally driven design approach. This is particularly evident in the
Rolex Learning Center, where site-specific data like movement patterns of people and
scenic viewpoints of the surrounding mountains inform the creation of an alternative
topography hovering above the existing one. In SANAA’s design, Rowe’s hierarchical
figure–ground relation dissolves into an interaction of two surfaces of equal value,
the figure–surface and the ground–surface, with the interaction described by a set of
geometric operations acting on the given data. The conceptual dissolution of figure
and ground within an operative design approach is also apparent in the Yokohama
26 2 Convergence

Port Terminal by Foreign Office Architects in the sense of Rowe, where both surfaces
are figure and ground at the same time.
In general, an operative design approach is dependent on the availability of data
for input [38]. This necessity forms the starting point for the so-called Layer Model
by McHarg that he introduced as a foundation for a design and planning theory
applied within landscape architecture [44]. In this model, each layer represents site-
specific data on one phenomenon like the occurrence of wildlife, plants, or soil types
as well as surficial and bedrock geology or surface and ground water hydrology. With
this collection of data, McHarg was aiming at the study of the interaction between
these layers as parts of the whole, thus focusing on a holistic understanding of the
interrelationships of natural and human systems.
For this, he superimposed the layers to show composite data and how each layer
of data relates to each other. But it is in this collage of layers of data where McHarg’s
approach of a holistic understanding of interrelationships falls short. The visual
layering of data provides only a superficial correlation between layers that is unable
to unravel deeper relational structures [40], a shortcoming apparent in the design of
Fresh Kills Park by Field Operations.
McHarg’s method of data collage is comparable to the layering in architectural
design strategies characteristic for Postmodernism and Deconstructivism that Greg
Lynn criticized as a “close analysis of contextual conditions from which they proceed
to evolve either a homogeneous or heterogenous urban fabric. … [None of them]
seems adequate as a model for contemporary architecture and urbanism” [42].
What McHarg was lacking is a computational design approach that presents a
new paradigm to architectural representation beyond the visual and motivates “a
fully integrated systemic approach ranging from bits, to codes to the structuring
of relationships” [40]. In a computational design, layered data is associated in an
explicit manner by means of transformations and geometric operations that describe
the interaction between data. This is illustrated in the renaturation project of the river
Aire, where operations capture the dynamics of the flowing water and its interaction
with the riverbed. The explicit availability of the rules of interactions enables the
design of a gridded flood plan that makes visible the geometric rules inherent in the
flow of water through the successive transformation of the grid.
In the project of river Aire, the structuring of relationships is primarily focused
on the interaction of data between the layers of surface hydrology and surficial
soil. Within this setting it can be differentiated between operations acting upon data
within a single layer, so-called horizontal operations, and operations acting upon
data between multiple layers that are vertical ones. Horizontal operations within the
layer of surface hydrology are defined by the relationships that capture the dynamics
inherent in flowing water. Vertical operations between the layers of surface hydrology
and surficial soil, on the other hand, are defined by the relationships that describe
the interplay of the flow of water with the geometry of the riverbed. Horizontal and
vertical operations make explicit the interrelationships inherent in layers of data and,
thus, can act as active agent in the design process.
Allen’s notion of field condition comprises this action of horizontal and vertical
operations in a quintessential way: horizontal relationships describe the flexibility
2.6 Contextual Figuration of Ground 27

within the organizational pattern of the field and vertical relationships the adaptation
of the field to external impacts. This interplay of internal organization and external
adaptation is the main driver in Allen’s initial study of the Jeffersonian grid as a
bottom-up local adaptation of a flexible grid to the given conditions on site.
It is this interplay of internal organization and external adaptation governed by
the horizontal and vertical operations that make McHarg’s initial goal of a planning
and design tool grounded in a holistic understanding of the interrelationships of
natural and human systems possible. The same is true for Rowe’s figure–ground
diagram with which he was aiming at opening up a dialogue between buildings and
their surroundings in order to overcome the object fixation of modern architecture.
Horizontal and vertical operations allow for a much deeper weaving of buildings
and surroundings into a coherent urban texture that reaches far beyond the visual
layering of Rowe’s initial design tool.
The operative interaction of layers of data of the natural and manmade environ-
ment is a weaving-together, a contextus as “joining together” in the truest sense and
defines a structurally grounded contextualism that overcomes the dualism of object
and neighborhood. It dissolves the difference between natural and manmade and
describes the computational framework for the contextual figuration of the ground
(Fig. 2.12).

Fig. 2.12 Context within a computational design approach as layered set of data that allows for
exploration of relationships by means of horizontal and vertical operations as foundation for a
process of figuration
28 2 Convergence

References

1. Allen S (1997) From object to field. AD Archit Des 67(5–6):24–31


2. Allen S (2010) Field conditions revisited. Long Island City, New York
3. Allen S, McQuade (2011) Landform building. Lars Müller Publishers, Baden
4. Auret H (2018) Christian Norberg-Schulz’s interpretation of Heidegger’s philosophy: care,
place and architecture. Routledge, London
5. Balmori D, Sanders J (2011) Groundwork: between architecture and landscape. Monacelli
Press, New York
6. Belogolovsky V (2016) Paulo Mendes da Rocha: “Architecture does not desire to be functional;
it wants to be opportune”. ArchDaily. http://www.archdaily.com. Accessed 3 July 2020
7. Berrizbeitia A (2016) On the limits of process: the case for precision in landscape. Lecture
notes. Harvard University Graduate School of Design, delivered 14 May 2016.
8. Betsky A (2002) Landscrapers. Thomas & Hudson, New York
9. Bollinger K, Grohmann M et al (2010) Das Rolex learning center der EPFL in Lausanne.
Beton- und Stahlbetonbau 105(4):248–259
10. Bucher A (2020) AIRE: the river and its double. J Landsc Archit 15(2):90–92
11. Corner J (1999) Recovering landscape: essays in contemporary landscape architecture.
Princeton Architectural Press, New York
12. Corner J (2010) The map-art, an interview with James Corner: some final thoughts. In: Amoroso
N (ed) The exposed city: mapping the urban invisibles. Routledge, London
13. Daglioglu EK (2015) The context debate: an archaeology. Archit Theory Rev 20(2):266–279
14. Dangermond J, Steinitz C (2015) The lab for computer graphics and spatial analysis (1965–
1991) and its legacy. The Center for Geographic Analysis, Cambridge
15. Deleuze G, Guatari F (1987) A thousand plateaus: capitalism and Schizophrenia. University
of Minnesota Press, Minneapolis
16. Descombes G, Descombes J, Cauwenberghe CV, Correnti V, Gerber F (eds) (2018) AIRE: the
river and its double. Park Books, Zurich
17. Deutsch R (2017) Convergence: the redesign of design. Wiley, London
18. Djokić V, Bojanić P (2017) Peter Eisenman: in dialogue with architects and philosophers.
Mimesis International, Milano
19. Dring M, Wall E (2015) Landscapes of variance: working the gap between design and nature. In:
Czechowski D, Hauck T, Hausladen G (eds) Revising green infrastructure: concepts between
nature and design. CRC Press, Boca Raton
20. Dutton G (1977) An extensible approach to imagery of gridded data. In: Proceedings of the
4th annual conference on computer graphics and interactive techniques SIGGRAPH 77. ACM
Press, San Jose, California, pp 159–169
21. Etymonline (2020) Online etymology dictionary. https://www.etymonline.com. Accessed 30
Apr 2020
22. Ferrater C (2006) Synchronizing geometry: landscape, architecture & construction. Actar,
Barcelona
23. Ferrater C, Ayala N (eds) (2015) OAB: office of architecture in Barcelona. Actar, New York
24. Gandelsonas M (1999) X-URBANISM: architecture and the American City. Princeton
Architecture Press, New York
25. Girot C (2011) Sigirino Depot-Switzerland: large artificial mound of excavation material. T
Topos-Eur Landsc Mag 74:73
26. Girot C (2013) Immanent landscape. Harvard design magazine. Landscape architecture’s core?,
pp 6–16
27. Girot C et al (2013) Topology: topical thoughts on the contemporary landscape. Jovis, Berlin
28. Girot C (2016) The course of landscape architecture: a history of our designs on the natural
world, from prehistory to the present. Thames & Hudson, London
29. Graves C (2020) The legacy of Colin Rowe and the figure/ground drawing. In Tice J, Hurtt S,
Latini AP (eds) The urban design legacy of Colin Rowe. ORO Editions, San Francisco
References 29

30. Hawkes D (2008) The environmental imagination: technics and poetics of the architectural
environment. Routledge, London
31. Hawkes D (2013) The environmental tradition: studies in the architecture of environment.
Routledge, London
32. Herrington S (2010) The nature of Ian McHarg’s science. Landsc J 29(1):1–20
33. Herrle P, Schmitz S (2009) Constructing identity in contemporary architecture: case studies
from the South. Lit Verlag, Berlin
34. Hilberseimer L (1944) The new city: principles of planning. Paul Theobald, Chicago
35. Jauslin D (2019) Landscape strategies in architecture. Dissertation, EPFL
36. Juel Clemmensen T (2014) The management of dissonance in nature restoration. J Landsc
Archit 9(2):54–63
37. Köhler D (2016) The mereological city: a reading of the works by Ludwig Hilberseimer.
Transcript Verlag, Bielefeld
38. Kotnik T (2010) Digital architectural design as exploration of computable functions. Int J
Archit Comput 8(1):1–16
39. Leatherbarrow D (2004) Topographical stories: studies in landscape and architecture. Univer-
sity of Pennsylvania Press, Philadelphia
40. Lorenzo-Eiroa P (2013) Form:In:Form—on the relationship between digital signifiers and
formal autonomy. In: Lorenzo-Eiroa P, Sprecher A (eds) Architecture in formation: on the
nature of information in digital architecture. Routledge, London
41. Lynch K (1960) The image of the city. MIT Press, Cambridge
42. Lynn G (1993) Architectural curvilinearity: the folded, the pliant and the supple. In Lynn G
(ed) Folding in architecture. Willey-Academy, London
43. Marshall C (2015) Pruitt-Igoe: the troubled high-rise that came to define urban America—a
history of cities in 50 buildings, day 21. The Guardian. http://www.theguardian.com. Accessed
24 Apr 2020
44. McHarg IL (1969) Design with nature. Doubleday, Garden City
45. Menges A (2008) Integral formation and materialization: computational form and material
gestalt. In: Kolarevic B, Klinger K (eds) Manufacturing material effects: rethinking design and
making in architecture. Routledge, London
46. Milheiro AV (2019) Retrospective: Paolo Mendes da Rocha. The architectural review. http://
www.architectural-review.com. Accessed 23 June 2020
47. Mostafavi M (2003) Landscape urbanism: a manual for the machinic landscape. AA
Publications, London
48. Nesbitt K (1997) Theorizing a new agenda for architecture: an anthology of Architectural
theory 1965–1995. Princeton Architectural Press, New York
49. Norberg-Schulz C (1982) Genius loci: towards a phenomenology of architecture. Rizzoli, New
York
50. Obiol C, Marques S (2011) An interview with Paulo Mendes da Rocha. Palimpsesto 3:33–41
51. Pollak L (2002) Fresh kills—sublime matters, pp 40–47
52. Rosenberg E (2019) Before and after. Both. The revitalization of the Aire River, Switzerland.
Landsc Archit Mag122–132
53. Rowe C, Koetter F (1978) Collage City. MIT Press, Cambridge
54. Ruby I, Ruby A (2005) Groundscapes: the rediscovery of the ground in contemporary
architecture. Editorial Gustavo Gili, Barcelona
55. Shepard DS (1984) Computer mapping: the SYMAP interpolation algorithm. In: Gaile GL,
Willmott CJ (eds) Spatial statistics and models. Springer Netherlands, Dordrecht
56. Steinitz C (2013) Beginnings of geodesign: a personal historical perspective. Research in
urbanism series. https://journals.open.tudelft.nl/rius/article/view/1366. Accessed 10 Nov 2020
57. Treib M (2016) Process and product: the “renaturalization” of the River Aire. In: Petschek P,
Siegrist D, Tschumi C (eds) Bridging the gap. Rapperswil, Switzerland
58. Venturi R (1966) Complexity and contradiction in architecture. Museum of Modern Art Press,
New York
59. Waldheim Ch (2006) The landscape urbanism reader. Princeton Architectural Press, New York
30 2 Convergence

60. Waldheim Ch (2011) The invention of GIS. The Harvard Gazette. https://news.harvard.edu/
gazette/story/2011/10/the-invention-of-gis/. Accessed 11 Nov 2020
61. Walliss J, Rahmann H (2016) Landscape architecture and digital technologies: re-
conceptualising design and making. Routledge
62. Will T (1988) Kontextualismus: Eine Stadt(um)baumethode. Baumeister 8:44–50
63. Wortham-Galvin B (2010) The Woof and the Warp of architecture: the figure-ground in urban
design. Footprint 4(7):59–74
Chapter 3
Patterns of Interaction

3.1 Topological Turn

Landscape Formation One, LF One, is a pavilion designed by Zaha Hadid as part of


the Landscaping and Gardening Exhibition held in Weil am Rhein, Germany in 1999.
Used primarily for exhibitions and for socializing, the building marks the entrance
to a redesigned landscape of an abandoned gravel quarry. Built a few years after
the completion of Hadid’s famous fire station for the campus of the Vitra company,
located only few kilometers away, LF One has received much less critical attention
despite its immediate relevance for a number of subsequent competition entries by
Hadid.
LF One is a test case, the built proof of concept of merging building and landscape,
an architecture as “purely synthetic, invented landscape that becomes as critical as
the real landscape. This artificial landscape replaces the natural rock formation.…
There is no longer a natural topography but cantilevers and ramps that become
the spectacle” [23]. For Hadid, the notion of landscape describes an organizational
concept that fosters “open, fluid, ambiguous organizations … to make a site much
more porous to allow for flows of any kind to move through it. Maybe not entirely
or indiscriminately, but maybe a selective acceleration of flows, or a branching of
flows into many alleys, which allow people to follow their trajectories, events to take
place, congregation, meeting places, sitting out or whatever. So these sites in the city
are no longer private territories but rather territories of the city” [23].
For Hadid, the notion of landscape embodies an urban concept of public porosity.
And it is this kind of porosity that is explored by Hadid in LF One for the first time in
a built project. In the Vitra fire station such porosity does not exist. Various cuts in the
volume allow for a transparency that visually connects the inside with the outside.
But the building remains a closed entity on site that does not invite the public. It is
still a private territory of exclusion, a limitation not only caused by the design itself
but also by its functionality as a fire station and its surrounding context as part of an
active production facility.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 31
P. Fricker and T. Kotnik, Patterns of Interaction,
SpringerBriefs in Architectural Design and Technology,
https://doi.org/10.1007/978-981-19-9083-0_3
32 3 Patterns of Interaction

Fig. 3.1 Organization of LF One’s building program into linear sequences of spaces (a); visible
in section as interacting spatial volumes (b); the bundle of spaces provides new paths into the site
through, over, and around the building (c) (Images b and c: Zaha Hadid Architects)

LF One, in contrast, is conceptualized as public territory, as an inviting gesture


into the site of the former quarry. The building gradually arises out of the ground like
an artificial rock formation, providing a point of orientation in the abandoned site:
a tabula rasa that is being structured by the emerging building. LF One is designed
as a bundle of linear sequences of spaces that engage the site like fibers defining
radiating directions of continuation (Fig. 3.1c). It is not only a building that reacts to
the existing field condition but actively acts as a force that transforms the surrounding
field itself by implied trajectories of the fiber bundle.
Each fiber relates to a sequence of spaces that are defined by the program of
the building: a series of offices and meeting rooms, an exhibition area, and a bistro
(Fig. 3.1a). The relative independency of these programmatic clusters and difference
in spatial requirements allows for flexibility. In section, this flexibility between the
programmatic sequences is articulated as shifted volumes resulting in a differentiated
roofscape, an artificial landscape that fosters movement over the building (Fig. 3.1b).
This roofscape is made accessible along the root of the bundle that emerges out of the
ground as a gently rising ramp inviting people to occupy the fifth façade as additional
ground. LF One demonstrates “the idea of multiplicity of grounds. The question [for
design] was how to create many grounds and thus strategically create many civic
public layers, to intensify civic activity” [23].
Hadid resolved this design question by allowing the architecture to become path-
like and the paths to become architecture-like. Such fluidity in the identity of archi-
tectural elements is captured in the concept of the fiber bundle and marks a shift
in Hadid’s design methodology. The design of the Vitra fire station was still based
on the placement of walls, of static primary elements with space as the residual in
between. The design of LF One, on the other hand, is based on spatial strands without
predefined shape. This requires a fluid geometric notion that is not encapsulating a
specific morphology but rather inherent spatial relationships.
3.1 Topological Turn 33

With LF One, Zaha Hadid moves from design thinking based on metric geometry,
or the shape of elements, to a design thinking grounded in topology, or the relationship
between elements. Topology defines the patterns of connectivity; in the case of LF
One, it is visualized as bundle of sequence of spaces (Fig. 3.1a). This bundle describes
the set of rules that governs the inner organization of the project and defines the
insertion points for external forces to interact, such as the various paths of movement
which intertwine with the spatial bundle (Fig. 3.1c).
The example of fiber bundles as a geometric design concept illustrates the topo-
logical shift in Zaha Hadid’s design thinking in the late 1990s. This shift may have
been triggered by the introduction of computational means into the design process
[43] that made topological design methods feasible and with it the idea of multiplicity
of grounds. In Hadid’s LF One, the architecture is not an added object but unfolds
from the existing ground condition. The topological gets topographic (Fig. 3.2) and
the architecture merges with its surroundings to create a new ground condition. The
fiber bundle, or unique pattern of connectivity, finds its form as an artificial rock
formation out of an interaction with existing topographic conditions. With this, LF
One illustrates how topological design thinking allows for adjusting an inner orga-
nization of an architectural design to external contextual situation in a flexible way.
This means the topological approach describes an adaptive process of mediation of
the generic of the architectural concept to the specific of the local condition.
Hadid further explored the architectural potential of a topological design strategy
like the fiber bundle as a strategy of territorialization [43] more systematically in
a number of subsequent competition entries. Following LF One, the fiber bundle
strategy was used in the design of the New Campus Center for the IIT in Chicago in
order to stitch a new pattern of movement into the regular urban grid. This competition
entry functioned almost as a diagram of the Maxxi Contemporary Arts Center in
Rome [23], where a bundle of buildings activates the idea of a multiplicity of grounds
within a complex urban surrounding.

Fig. 3.2 LF One is conceptualized as an artificial rock formation that guides the visitor into the aban-
doned gravel quarry (a). It unfolds out of the ground as part of the existing topography (b) (Image:
Zaha Hadid Architects)
34 3 Patterns of Interaction

As a test field of new ideas, Hadid’s LF One illustrates a weaving of the building
and surrounding path systems into a coherent texture, a literal contextus as “joining
together” a newly established context of multiple grounds. LF One truly is a figuration
of the ground, an architecture as a landscape formation, as the new ground itself.
The resulting spatial experience is essentially defined by the movement patterns on
topographic data made operative by the geometric logic of the fiber bundle. In other
words, the design as figuration of the ground is primarily a manipulation of the
ground by defining layers of data in the natural and manmade environment.

3.2 Topological Design Thinking

With the move towards topological design, Zaha Hadid spearheaded a development
that Greg Lynn described as a turning point in contemporary architectural theory in
his publication Folding in Architecture [30]. Lynn’s publication was a catalyst for a
wave of change that marked the turn of the millennium, when the avantgarde that
evolved out of it was regarded as “the quintessential architectural embodiment of the
new digital technologies that were booming at that time” [11].
For Lynn, the common thread in the work of this avantgarde was a “desire for
architectural complexity in both composition and construction” grounded in topolog-
ical thinking [15]. The topological approach in architecture was seen as a conscious
attempt to move beyond Postmodernism and Deconstructivism. Both architectural
reactions are based on “a close analysis of contextual conditions from which they
proceed to evolve either a homogeneous or heterogeneous urban fabric. … [None
of them] seems adequate as a model for contemporary architecture and urbanism.
Instead, an alternative smoothness is being formulated that may escape these dialec-
tically opposed strategies” [29]. It is here where Lynn turns to topology as “a flexible
system for the organization of disparate elements within continuous spaces” [29].
Hadid’s use of the fiber bundle in the design of LF One illustrates the flexible
system of organization very well. And the affordability of computers and the avail-
ability of software was key in making topological methods of design more acces-
sible in practice over the following two decades. But topological design thinking in
architecture did not originate in the late 1990s.
Already in the 1950s topology and other mathematical sub-disciplines played a
major role in developing a design methodology at the Hochschule für Gestaltung
(HfG) in Ulm, the Ulm School of Design. Together with the Bauhaus, the HfG is
widely considered as one of the most important design schools of the twentieth
century [19] with a reputation of being a “citadel of methodolatry” [31]. The HfG
Ulm had no access to computers but with their theorization and pedagogy they were
opening up the new field of computer science to the discipline of architecture [37].
An important characteristic of the curriculum of the HfG was the emphasis on
the application of both scientific knowledge and scientific methods in the design
process. Especially emerging scientific disciplines like system theory, cybernetics,
and information theory were seen as essential for an understanding of a machine
3.2 Topological Design Thinking 35

theory with the aim to “furnish the designer with a vocabulary of terms to analyze
and to quantify the structural relationships” [31].
Topology was of particular value in architecture as well as town and regional
planning where complex circulation problems must be solved because “topology
provides the designer with yet another approach to his problems. With the aid of
topology, he can discover that they are not only dimensional, form and position
problems but also problems of organization, continuity, and neighborhood. In other
words topology encourages the designer not to approach the world of technical
objects solely in metric terms but also in non-metric terms” [31].
Topological notions infused the whole pedagogy of the HfG and made it into
a nexus and promoter of topological developments in European design [49]. At the
HfG, topology was seen as a geometric language that allows for the study of the struc-
ture of spatial organization and the inherent part-to-whole relationships. Systematic
exploration and communication of these relationships by means of graphs, grids,
lattices, and networks (Fig. 3.3) were essential parts of the so-called visual method-
ology already at foundation courses at the HfG [20]. With its pedagogy, the HfG repre-
sented one of the most thorough early attempts to inject topological ideas directly
into the critical conversation of design [49].
A similar embracement of topological methods appeared at the University of
Cambridge in an attempt to build up a scientific foundation for the discipline of
architecture. From the late 1950s through the 1970s, much of the Cambridge depart-
ment had been explicitly committed to establishing architecture as a science grounded
in a body of quantifiable knowledge through research [25]. Many of the influential

Fig. 3.3 Examples of exercises from the foundation course at HfG: a studies of polyhedra that
completely fill the space based on regular and semi-regular lattices, and b exploration of possible
developments of the street network of Zürich 1955–1958 and corresponding representation as graph
[20]
36 3 Patterns of Interaction

theorists like the operations researcher Horst Rittel were convinced that topology was
an essential basis for a truly modern mathematical theory of architecture [49]. And
the work of the Cambridge architects Lionel March and Philip Steadman can be seen
as proof of Rittel’s proposition. They applied graph networks to define topological
invariants of floor plans which allowed them to discuss the equivalence of layouts
through the corresponding equivalence of underlying network structures [45].
It was in this atmosphere of rethinking architectural principles and design method-
ologies that Christopher Alexander studied architecture and mathematics at the
University of Cambridge. Strongly influenced by the “drift of the Cambridge work
away from buildings as material, social, and symbolic construction and towards
the abstraction of geometry and mathematics” [25]. Alexander went onto explore
this direction of thinking more thoroughly as a student in the new PhD program in
architecture at Harvard University where he completed his dissertation in 1962 and
published it in 1964 as Notes on the Synthesis of Form [2].
In his doctoral research, Alexander views the design process as the complex
interaction of a large set of diverse requirements that need to be met. By formal-
izing these interactions into a network of relationships between quantifiable and
non-quantifiable variables, Alexander is able to abstract the design problem by inves-
tigating an equivalent graph. By subdividing this graph into subgraphs (Fig. 3.4a) it is
possible of subdividing the design problem into subproblems that are interlinked. As
a consequence, each design problem can be understood as hierarchical composition
of subproblems “whose structural hierarchy is the exact counterpart of the functional
hierarchy established during the analysis of the problem” [2].
Every subgraph therefore can be perceived as diagrammatic representation of a
common architectural problem. For Alexander it is this result of his dissertation that
stands out as the key finding—the idea of the diagram. In retrospect he wrote in the
foreword to the paperback edition of Notes on the Synthesis of Form:
These diagrams, which, in my recent work, I have been calling pattern, are the key to the
process of creating form. … The idea of a diagram, or pattern, is very simple. It is an abstract
pattern of physical relationships, which resolves a small system of interacting and conflicting
forces, and is independent of all other forces, and of all other possible diagrams. The idea

Fig. 3.4 Alexander’s Notes on the Synthesis of Form introduces a methodology of translation of
design problems into interlinked graphs (a), and this method is illustrated in the epilogue of his
dissertation using the design of a village in India as example (b) [2]
3.2 Topological Design Thinking 37

that it is possible to create such abstract relationships one at a time, and to create designs,
which are whole by fusing these relationships – this amazingly simple idea is, for me, the
most important discovery of the book.

Alexander continued and intensified his research into topological diagrams as


foundation for architectural design at his own Center for Environmental Studies at
Berkeley University and verified the applicability of patterns during an experimental
approach to campus planning at the University of Oregon. This resulted in the publi-
cation of A Pattern Language [4], a collection of 253 patterns presented as practical
sourcebook reaching in scope from regional planning down to construction details
aiming at the design of “towns and neighborhoods, houses, gardens, and rooms”.
For Alexander, these patterns describe design considerations of archetypical value
on all scales of architectural intervention. They are of timeless validity and define the
core of any human-centered environment because patterns are “so deep, so deeply
rooted in the nature of things, that it seems likely that they will be a part of human
nature, and human action, as much in five hundred years, as they are today” [4]. One
of these archetypical patterns deeply rooted in human nature is the need for “access
to the countryside, experience of open fields, and agriculture, access to wild plants
and birds and animals” [4]. Based on the assumption that the open country should be
within 10 min of walking distance, the pattern defines an urban typology of deeply
interlocked stretches of urban space and countryside (Fig. 3.5a).
This pattern of so-called city-country fingers illustrates some of the confusion
inherent to A Pattern Language, which has weakened its reception and resulted in
rejection of Alexander’s work by many contemporary architects and scholars since
[14]. With the pattern Alexander raises awareness of the importance of green spaces
for the well-being of city residents, a topic that has been studied in depth and more
scientific rigor only in the twenty-first century [48]. But at the same time the forward-
looking understanding of urban planning is counteracted by the practical attitude of
the book as an easy-to-use guide for the layman. This has resulted in obscuring the
general topological diagram with exemplary specificity.
In the case of the pattern of the city-country fingers, the understanding of the
importance of the topology of the boundary between the manmade and the natural
environment has been overwritten by an articulation of this boundary within a
predominantly agrarian society of small-scale farmers. The pattern of city-country
fingers, therefore, has to be understood as a specific example within a particular
context and not as the general topological diagram of the relationship between the
manmade and the natural environment.
This differentiation is often ignored by readers of A Pattern Language but pointed
out by Alexander in his introduction. Topology as language of structural relationships
is abstract and because of this “it is not possible to solve the stated problem properly,
without shaping the environment in one way or another according to the pattern
that we have given” [4]. Every diagram is inevitably a specific actualization of the
inherent structure, a representation of the general by the specific. And this general,
captured in the pattern of the city-country fingers, is the insight that urban space needs
38 3 Patterns of Interaction

Fig. 3.5 Alexander’s pattern like the city-country fingers (a), describe a topological design
approaches that has been refined in recent years by strategies of operative explication (b), and
structural differentiation (c) [2, 22, 24]

to maintain an openness towards its surrounding, a permeable strategy of territorial


occupation.
Such topological explorations of strategies of territorial occupation have reap-
peared over the past two decades, caused by a topological turn in architecture.
Alexander’s pattern of city-country fingers has resurfaced as Rurbanizing, an environ-
mental logic in the work of Vincent Guallart (Fig. 3.5b). Rurbanizing is concerned
with the “creation of an urban edge in a city, maintaining an open structure that
connects with the natural networks of the environments that penetrate the city” [22].
Guallart avoids the communicative difficulties of Alexander by making the operative
process of construction of the urban edge as an explicit additional geometric logic like
topographying, reconnecting, or geomorphosis. This means the topological diagram
and its specific actualization are kept apart and not morphed into one single pattern
as in the case of Alexander.
A different approach is visible in the work of Joachim Huber on urban topologies
[24]. Here, Alexander’s topological diagram is not supplemented by an operative
explication. Rather, the topological diagram itself is continuously differentiated by
the application of methods from the mathematical field of topology (Fig. 3.5c).
3.3 Patterns That Connect 39

Notions like homotopy, homology, or fiber bundles are used as tools in a struc-
tural exploration of urban phenomena and the boundary of urban spaces. With
this, Alexander’s pattern of city-country fingers unfolds into a multitude of urban
configurations and boundary conditions.

3.3 Patterns That Connect

The discussion shows why Alexander’s A Pattern Language received a lot of criticism
from architects and scholars and, at the same time, “could very well be the most read
architectural treatise of all time, yet in the architecture schools I know, it is as if this
book did not exist” [41]. The 253 patterns describe a particular architectural language
that has its origin in the study of vernacular architecture and medieval cities, a fact
that has obscured the discussion of Alexander’s ideas [14].
But as demonstrated, A Pattern Language should not be understood as a guidebook
for application, despite its intention, but rather as an extensive example that illustrates
a design methodology based on topological diagrams. A methodology of design as
problem-solving as already sketched out by Alexander in Notes on the Synthesis of
Form. In other words, A Pattern Language makes plain a specific logic of problem-
solving.
This logic displays some similarity with the work of political scientist Herbert A.
Simon and his article The Architecture of Complexity [44]. Simon was searching for
principles of organization shared by complex systems studied within the emerging
field of general systems theory. He noticed the tendency in complex systems to
form so-called nearly decomposable hierarchies characterized by weak interactions
between hierarchical units. This allows systems to be broken into sub-systems that
can then be explored in a more comprehensible way, because “subparts belonging to
different parts only interact in an aggregative fashion—the details of their interaction
can be ignored” [44].
This is an observation that Alexander shared in principle and led him to the
definition of patterns as independent diagrams of interaction in Notes on the Synthesis
of Form (Fig. 3.4). As a consequence, Alexander viewed architectural design as
“diagrammatoidal reasoning and problem solving. The patterns, then, are both the
means and the rules of this kind of reasoning: abstracted solutions which will become
concrete whenever they are applied to a specific context” [9].
Thus, A Pattern Language can be perceived as a formal language of indepen-
dent but hierarchically organized patterns of interaction. It is this perspective that
made software developers pick up on Alexander’s ideas in the late 1980s. The
notion of patterns heavily influenced the development of object-oriented program-
ming languages like Smalltalk or C++ [28]. Technologies like Wiki were a direct
outgrowth of Alexander’s pattern language [13]. Around the turn of the millennium,
such digitally encoded patterns found their way back into architecture by means of
topological modeling tools [50].
40 3 Patterns of Interaction

Accordingly, the topological turn in architecture [15] can be understood as a turn


towards design patterns, or an unconscious activation of Alexander’s conception of
architectural design mediated by computational tools and scripted codes. Each of
Alexander’s design patterns is a diagrammatic expression of the interaction with the
multidimensional cultural, political, social, and physical context of architecture like
illustrated by the fiber bundle of Zaha Hadid’s LF One, which aligns the programmatic
organization of the building with the adjoining street network and topography.
The topological pattern is the study of the place, the analysis situs as practiced by
Leonhard Euler in his solution of the problem of the Seven Bridges of Königsberg
in 1736, which prefigured the idea of topology [16]. That is, the topological pattern
describes the use of contextual information and its interaction that is the horizontal
and vertical interaction of data within a digital design approach (Fig. 2.12). It is the
pattern that connects layers of data.
But the pattern is in its interaction not limited to the layers of data it operates
on. Each topological pattern describes one element in a nearly decomposable hier-
archy, as Herbert Simon has described it. In line with Simon’s intention, Christopher
Alexander explored the part-to-whole relationship within pattern languages. Contrary
to Simon, Alexander focused more on the property of interconnectivity between
nearly decomposable patterns and the consequent network aspect [13]. Such inter-
connectivity is of great relevance for the creation of a web-like organization between
patterns and this observation ultimately resulted in the publication of the widely cited
essay A City is Not a Tree [3].
For each of the 253 patterns in A Pattern Language, this web-like connectivity is
articulated as a list of related patterns describing in its totality an adjacency matrix of
the pattern network, a matrix that illustrates not only that cities are not structured like
trees. In addition, the adjacency matrix also shows that the well-known architectural
hierarchies of design thinking like the region, the city, the neighborhood, the city
block, or the individual building are not decomposable hierarchies of independent
intervention, because “no pattern is an isolated entity. Each pattern can exist in the
world only to the extent that is supported by other patterns: the large patterns in
which it is embedded, the patterns of the same size that surround it, and the smaller
patterns which are embedded in it” [4].
All traditional levels of architectural thinking interact with each other and, ulti-
mately, this endless nesting of patterns places the entirety of the manmade world in
relation to the natural world we live in and the patterns that govern it. “You cannot
merely build that thing in isolation … the thing which you make takes it place in the
web of nature, as you make it” [4]. The topological perspective reveals the organiza-
tional schema of architecture, while the topological diagrams or patterns relate back
to the natural environment and ground all architectural organization. Architecture
needs to design with nature.
This conclusion reformulates the findings of Ian McHarg in his pioneering book
Design with Nature [32]. This publication not only signaled the high-water mark
of the ecological movement in the United States but also strongly influenced the
development of landscape architecture and environmental planning for the past half
century [18]. McHarg argues that every architectural intervention would benefit from
3.3 Patterns That Connect 41

being designed with regard to both the inherent laws of nature and the characteristics
of the specific site and surrounding landscape.
This means the natural world is not perceived anymore as the other in opposition to
the manmade. Rather it permeates the manmade environment on all levels, resulting in
an overlay of “two systems within an metropolitan region—one the pattern of natural
processes preserved in open space, the other the pattern of urban development” [32].
Christopher Alexander articulated this basic conception of interlocking urban design
as a pattern of city-country fingers (Fig. 3.5). Contrary to Alexander’s ambition of
extracting abstract principles, McHarg was primarily interested in the complexity of
the specific which he analyzed by means of layers of data, a design method which
predated the invention of geographic information systems (GIS) and other mapping
technologies. McHarg’s pattern of city-country fingers, therefore, is a pattern of
high-resolution differentiation (Fig. 3.6), which celebrates the manifestation of the
pattern in reality rather than the illustration of an underlying principal or mechanism
of action.
Topological patterns provide an epistemological perspective of the world. In
general, patterns operate according to an aesthetic logic, one that is based on “recog-
nition and empathy” rather than rationality as Gregory Bateson has pointed out in his
study of “patterns that connect” [8]. They provide a conceptual framework: patterns
are a mental construction of “a picture of how the world is joined together” [8].
As diagrams, they provide orientation, new insights, and allow foster productivity,
because “there is a deep and important underlying structural correspondence between
the pattern of a problem and the process of designing a physical form which answers

Fig. 3.6 Summary maps of water and land features for part of the metropolitan region of
Philadelphia, which illustrates the permeation of the two systems of the natural and the urban
[32]
42 3 Patterns of Interaction

that problem” [8]. This structural correspondence not only links the diagrams of
Alexander and McHarg but also links architectural design with processes in nature.

3.4 Organized Matter

Laws of nature are pervasive and cannot be ignored in the design of the built environ-
ment without permanent additional effort, an effort that has driven the development
of architecture for thousands of years. Architecture has been shaped by the access to
energy and resources: from human labor to fire and fossil fuel [10]. Architecture, from
the scale of the building up to the scale of a metropolitan region, can be understood as
“a material organization that regulates and brings order to energy flows, and, simul-
taneously and inseparably, as an energetic organization that stabilizes and maintains
material forms” [17]. This means matter and energy are closely linked together, and
all buildings, cities, and landscapes—all organized matter—are subject to permanent
deterioration and in need of a constant supply of materials and energy in support of its
own reconstruction and stabilization. Matter and energy are the fundamental building
blocks of both, our manmade and our natural, environments.
Throughout history, however, architectural design thinking has primarily been
focused on the role of matter in architecture with the topic of energy routinely left
completely latent [34]. “Above all, architecture is matter that is arranged at a large
scale for the purpose of resisting gravity and the elements as well as for beauty and
use … [with] a deep resonance between the matter that architecture organizes and
the cognition of the mind” [39]. Over time, this understanding has promoted a highly
distinguished discourse on the difference between matter, materials, and materiality
and their role in architectural design. At the same time, the negligence of energy has
resulted in a built environment that is responsible for around one-third of CO2 global
emissions caused by the burning of fossil fuel for the construction, maintenance, and
operation of buildings.
The lack of concern for environmental issues in design thinking is especially
apparent in the fundamental alteration of the design of architectural envelopes after
World War II. The aim of controlled indoor climate displaced the traditional role of
the envelope of regulating heat gain, ventilation, and natural lighting for that of an
isolating paradigm aimed to separate interior from exterior [1]. The energetic isola-
tion of the building ironically had its conceptual foundation in the cooling industry
[35] and was achieved by introducing a multilayer envelope with minimum exposed
surface area and maximum insulation. With this, the topic of energy is treated as an
add-on to the design, as a supplementary technical necessity.
But energy is deeply embedded in the organization of matter. Matter is but captured
energy [34]. Like energy, matter is conserved. It is never created nor destroyed: it
cycles through our world. By the Law of Conservation of Mass, matter only changes
form. Any building, any city, and any landscape is “a set of vibrating molecular
lattices; an accumulation of molecular processes that eventuate in a form that main-
tains an organization for a certain duration. The bonds of these molecular lattices
3.4 Organized Matter 43

that compose matter itself are fundamentally energetic. Captured energy is the only
thing that maintains or alters matter’s bonds” [34]. Such deep entanglement of matter
and energy becomes particularly visible in seemingly spontaneous natural processes
of formations like in hydrodynamics, meteorology, geology, or metabolizing cyto-
plasmic extracts [27]. Modern thermodynamics shows how complex organizations
of matter come into being, expand, and increase their complexity by their exposure
to energy flows, capturing energy from their environment and dissipating it again
[42].
Energy flows are defined by gradients, differences of, for example, temperature,
pressure, or chemical concentration across a certain distance. Following thermody-
namic principles, these gradients tend to be eliminated by means of self-organization
of matter [42]. Processes of self-organization in thermodynamic systems (Fig. 3.7)
were first studied in depth by Ilya Prigogine, and he showed how gradients of energy
in dissipative systems can cause patterns or ordered organizations of matter to emerge
[40].
Matter is but captured energy. The nature of matter becomes visible in material
organization as residual evidence of the flowing of energy. Such an understanding
of formation processes in nature as the capturing and dissipating of energy from the
environment was first formulated by D’Arcy Wentworth Thompson, a pioneer of
mathematical biology, already several decades before Prigogine’s groundbreaking
work. In his widely admired book On Growth and Form, a study of morphogenesis
in plants and animals first published in 1917, Thompson pointed to the importance
of physical forces in the development of shapes in nature: “The form of an object
is a ‘diagram of forces’, in this sense we can deduce the forces that are acting or

Fig. 3.7 A dissipative system is an open thermodynamic system, which operates far from an equi-
librium. Such a system is characterized by the spontaneous formation of complex structures like
in the Belousov–Zhabotinsky reaction (left), an oscillating reaction with periodic variations in the
concentration of the reaction intermediators and catalysts. The Rayleigh–Bénard convection (right)
is occurring in a layer of fluids heated from below. The fluid develops a regular pattern of so-called
Bérnard cells based on the variation of density caused by the heated bottom of the fluid (Image:
Santa Fe Institute and University of Iowa, Instructional Resources)
44 3 Patterns of Interaction

Fig. 3.8 For the study of the impact of physical forces on the development of morphology, D’Arcy
Thompson used various diagrams like a the bending moment diagram for an understanding of the
load distribution within a skeleton or b diagrams of transformation for the impact of environmental
forces on bodies [46]

have acted upon it” [46]. Thompson argued that biological form is not necessarily
the result of a selection process but rather is inevitable, as the shape is dictated by
physical and chemical forces (Fig. 3.8). With this, growth is perceived as the process
of arranging matter according to the flow of energy. Form is neither governed by
purpose, nor does form follow function. Form follows energy.
At the time of publication, Thompson’s On Growth and Form was seen as a critique
of the Darwinian paradigm. But Thompson did not reject the theory of Charles
Darwin. Rather he argued with his research that the process of natural selection is
framed, or rather guided by rules of physics and chemistry [7]. Moreover, Thompson
was interested in the mechanism of the development of form meaning an operative
understanding of the patterns in nature. This is an answer to something the theory of
natural selection does not provide. Natural selection as the main driver of evolution
only acts on the genome, the hereditary information of an organism. But most genes
only carry instructions for the building of proteins. Therefore, the effect of genes
depends on the physical and chemical details of the biochemical process that the
protein undergoes. That is, genes only provide a blueprint for proteins, but the unique
topological pattern unfolds only under the action of separate physical and chemical
forces.
A topological pattern provides a principal schema for the arrangement of matter,
and the flow of energy adapts to the environmental condition. Like in Thompson’s
study of fishes (Fig. 3.8b) where the basic body plan is adjusted by the action of
geometric transformations of the underlying coordinate system caused by a “system
of forces [that] has been at work” [46]. This implies that the flow of energy defines
a field condition, “a triggering effect of organization on a whole set of elements, a
horizontal phenomena of figuration of the whole out of the interaction of smaller
individual parts” as Stan [5] introduced it (compare Chap. 2). As in the case of
3.4 Organized Matter 45

Allen, Thompson’s focus is on the operation that reveals how forces are acting upon
the topological pattern. This is because “skeletons are then not seen as arbitrary
structures molded this way and that by natural selection, but as constructions” [7]. It
is the operativeness that offers a deeper understanding of the underlying causes and
with it the possibility of a conscious adaptation in design.
The goal of operative activation of topological patterns is true for computationally
driven design as well. This may sound outdated, as computational tools and data-
driven approaches offer the possibility of simulating and predicting the interplay of
flows of energy and matter with increasingly unfathomable precision. But from a
design perspective, such use of computational power only provides an image, or a
static evaluation, as feedback. This kind of design approach resembles the attempts
of György Kepes to bridge science and art by visual readings of photographs from
science [26]. The gestalt-related reading had its origin in art education at the Bauhaus,
especially in the work of Laszlo Moholy-Nagy.
According to Kepes “images are the starting point of all our thinking and feeling
… Through images we participate in the world” [26]. In the 1930s and 40s, new
technologies provided a vast range of novel images of nature on both macroscopic
and microscopic scales that attracted him. In 1951, Kepes organized an exhibition
titled New Landscape in Art and Science that explored affinities among visual arts
and recent scientific visualizations with the aim that the natural patterns of organi-
zation made visible by photography would “clarify relations of order, continuity and
direction in the emergence, growth and disappearance of nature’s forms” [26]. But
most of the images that Kepes presented were images of homeostasis or processes in
equilibrium that no longer speak of the underlying dynamics of the generation and
the inherent interplay of energy and matter.
These photographs were very similar to computer-generated simulations or the
production of neural networks, both contemporary images of homeostasis. The
hidden operation or the black-box character of such design approaches resembles
very much D’Arcy Thompson’s critique on the Darwinian paradigm of hidden selec-
tive forces and its lack of insight into the actual mechanisms of form generation.
That is why the operative activation of topological patterns is also of importance for
computationally driven designs, because it focuses attention on the utilization of the
laws of nature within architectural design.
This understanding is what defines Frei Otto’s main motivation in the exploration
of principles of construction in nature, first at his interdisciplinary research group
Biology and Building at the Technical University in Berlin and from 1964 onwards at
the Institute of Light Construction (IL) of the University of Stuttgart. Otto’s mission
in architecture was to be in harmony with nature [36]. This fostered his research
into natural design because “we can study nature so we can be part of earth” [12].
He believed that the built environment must conform to the laws of nature. For this,
Otto worked like no other twentieth-century architect on investigating the emer-
gence of form in nature and developing a new form of light and natural, adaptable
and changeable building from an understanding that flowed from his research [36].
This research focused on so-called natural constructions based on processes of self-
organization and economic principles in nature with the aim of establishing rational
46 3 Patterns of Interaction

form-finding processes following natural laws. For this, he conducted numerous


experiments like the well-known soap film experiments on minimal surfaces [47]. But
he also studied organizational patterns in nature by collecting series of pictures very
much like Kepes did in his exhibition New Landscape in Art and Science [38]. Going
beyond Kepes’ interest in visual comparison, Frei Otto used these images to classify
natural phenomena and subsequently speculate on possible generative processes with
the goal of making the underlying principles of organization accessible for design
(Fig. 3.9).
Otto’s move from the image of a perceived pattern to the underlying rules of orga-
nization clearly demonstrates the understanding of patterns as topological schemas.
This differs very much from its use in contemporary discourse where the notion of
pattern points towards variations of regularity like, for example, in the context of
surface tessellations [21] or parametric variations of component systems [6]. But the
notion of patterns is more comprehensive: it is an operative description of relation-
ships between elements that can adapt flexibly to contextual changes. “Patterns are
synonymous with processes; they are indications of the forces and interactions that
created them” [33]. Patterns are the primary means by which the rise in ecological
consciousness has been expressed in design thinking and the methods determining

Fig. 3.9 Comparison of natural and manmade path systems (a) motivate the exploration into path
systems and the development of possible generative processes of networks with excess length
(b) [38]
References 47

these patterns ingeniously capture the relationships that are perceived in the world
[33].
Patterns offer an understanding of nature as a complex collection of environmental
operations that the built environment has to support or to substitute. By viewing
architecture, cities, and landscapes no longer as a collection of independent objects
of attraction but rather as an interacting network of relationships modulating local
flows of energy, the manmade becomes a symbiotic part of the natural environment,
making a more sustainable approach towards urbanization possible.
Patterns are not about simulating the world we live in nor a mere technological
mimicry of nature. Rather a creative investigation into natural patterns of interaction
and regular stasis can become a driver for design interventions.
With such an understanding, pattern-based architectural design intentionally shifts
away from the design of objects and the consumption of environmental resources
towards the design of interrelationships, of interaction with the environment. The
design of architecture, of cities, of landscapes and territories is about enhancing of
environments, it is about building environments.

References

1. Ábalos I (2017) Thermodynamic materialism. In García-Germán I (ed) Thermodynamic inter-


actions: an architectural exploration into physiological, material, territorial atmospheres. Actar,
Barcelona
2. Alexander C (1964) Notes on the synthesis of form. Harvard University Press, Cambridge
3. Alexander C (1965) A city is not a tree. Archit Forum 122(1):58–62
4. Alexander C (1977) A pattern language: towns, buildings, Construction. Oxford University
Press, New York
5. Allen S (1997) From object to field. AD Archit Des 67(5–6):24–31
6. Andersen P, Salomon D (2010) The architecture of patterns. Norton Company, New York
7. Ball P (1999) The self-made tapestry: pattern formation in nature. Oxford University Press,
Oxford
8. Bateson G (1979) Mind and nature: a necessary unity. Dutton, New York
9. Bauer M (2016) Pattern language and space syntax: Alexander, Chomsky, Pierce and Wittgen-
stein. In Krämer S, Ljungberg C (eds) Thinking with diagrams: the semiotic basis of human
cognition. De Gruyter, Berlin
10. Calder B (2021) Architecture: from prehistory to climate emergency. Penguin Books, London
11. Carpo M (2004) Ten years of folding. In Lynn G (ed) Folding in architecture, revised edition.
Wiley, London
12. Chiu S (2017) Frei Otto: spanning the future. http://www.freiottofilm.com. Accessed 27 July
2022
13. Cunningham W, Mehaffy M (2013) Wiki as pattern language. In: Proceedings of IEEE computer
society conference on computer vision and pattern recognition, vol 1, pp 32–47
14. Dawes MJ, Ostwald MJ (2017) Christopher Alexander’s A pattern language: analysing,
mapping and classifying the critical response. City Territ Archit 4:17
15. Di Christina G (2001) Architecture and science. Wiley, London
16. Euler L (1953) Leonhard Euler and the Koenigsberg Bridges. Sci Am 189(1):66–72
17. Fernández-Galiano L (2000) Fire and memory: on architecture and energy. MIT Press,
Cambridge
48 3 Patterns of Interaction

18. Fleming B (2019) Design with nature, Ian McHarg’s ideas still define landscape architec-
ture. Metropolis. https://metropolismag.com/viewpoints/mcharg-design-with-nature-50th-ann
iversary. Accessed 14 July 2022
19. Frampton K (1973) Apropos Ulm: curriculum and critical theory. Opposition 3:17–36
20. Fröshaug A (1959) Visual methodology. Ulm 4 J Hochschule für Gestaltung 4:57–68
21. Garcia M (2009) Patterns of architecture. AD Archit Des 79(6)
22. Guallart V (2009) Geologics: geography, bits and architecture. Actar, Barcelona
23. Hadid Z (2001) Landscape as plan. El Croquis 103. El Croquis Editorial, Madrid
24. Huber J (2002) Urbane Topologie: Architektur der randlosen Stadt. Universitätsverlag der
Bauhaus-Universität, Weimar
25. Keller S (2017) Automatic architecture: motivating form after modernism. University of
Chicago Press, Chicago
26. Kepes G (1956) The new landscape in art and science. Paul Theobald & Company, Chicago
27. Lamprecht I, Zotin AI (1988) Thermodynamics and pattern formation in biology. De Gruyter,
Berlin
28. Lea D (1997) Christopher Alexander: an introduction for object-oriented designers. http://
www.patternlanguage.com/bios/douglea.htm. Accessed 7 July 2022
29. Lynn G (1988) Folds, bodies & blobs: collected essays. La Lettre volée, Brussels
30. Lynn G (1993) Architectural curvilinearity: the folded, the pliant and the supple. In Lynn G
(ed) Folding in architecture. Willey-Academy, London
31. Maldonado T, Bonsiepe G (1964) Science and design. Ulm 10/11 J Hochschule für Gestaltung
10/11:10–29
32. McHarg I (1969) Design with nature. Natural History Press, New York
33. M’Closky K, VanDerSys K (2017) Dynamic pattern: visualizing landscapes in a digital age.
Routledge, London
34. Moe K (2011) Matter is but captured energy. In Borden GP, Meredith M (eds) Matter: material
processes in architectural production. Routledge, London
35. Moe K (2013) Insulating North America. J Constr Hist 27(1):87–106
36. Nerdinger W (2005) Frei Otto. Complete works: lightweight construction—natural design.
Birkhäuser, Basel
37. Neves I, Rocha J, Duarte J (2014) Computational design research in architecture: the legacy
of the Hochschule für Gestaltung, Ulm. Int J Archit Comput 12(1):1–25
38. Otto F (2009) Occupying and connecting: thoughts on territories and spheres of influence with
particular reference to human, Settlement. Axel Menges, Stuttgart
39. Picon A (2021) The materiality of architecture. University of Minnesota Press, Minneapolis
40. Prigogine N (1977) Self-organization in nonequilibrium systems: from dissipative structures
to order through fluctuations. Wiley, New York
41. Saunders W (2002) A pattern language by Christopher Alexander. Harv Des Mag 16:74–78
42. Schneider ED, Sagan D (2006) Into the cool: energy flow, thermodynamics, and life. University
of Chicago Press, Chicago
43. Schumacher P (2004) Digital Hadid: landscapes in motion. Birkhäuser, Basel
44. Simon H (1962) The architecture of complexity. Proc Am Philos Soc 106(6):467–482
45. Steadman JP (1983) Architectural morphology: an introduction to the geometry of building
plans. Pion, London
46. Thompson DW (1917) On growth and form. University Press, Cambridge
47. Vrachliotis G (2017) Thinking by modeling. Spector Books, Leipzig
48. WHO (2016) Urban green spaces and health: a review of evidence. Regional Office for Europe,
Copenhagen
49. Witt A (2022) Formulations: architecture, mathematics, culture. MIT Press, Cambridge
50. Yukita S (2002) Design patterns for topological modeling. 1st International Symposium on
Cyber Worlds, Tokyo, pp 455–464
Chapter 4
Computing Land-Scapes

4.1 Performative Patterns

The ongoing process of global urbanization not only affects the social, economic,
political, and cultural aspects of living, but has also caused an unpredictable impact
on our environment. Traditionally, urban life has been perceived as an escape from
natural conditions, the urban in opposition to nature. We are increasingly becoming
aware that urban development has to be understood as a development within and in
interaction with nature. This requires a rethinking of the design of our future cities.
The urban is understood as generated “land-scape”, as prosthetic nature based upon
a computational framework for the contextual figuration of the ground, as opposed to
the discussion in Sect. 2.6, which introduced architecture and landscape architecture
as a process-oriented transformation of given conditions to create new environments.
The built space should be a ground that is shaped in order to expand the inherent
logic and functionality of nature and challenges the convention of densification strate-
gies enhanced by ecosystem services. Landscape and natural processes are funda-
mental transitional concepts that operate on fields and boundary conditions. Reading
the urban fabric as the product of dynamic network systems with different contextual
layers allows a systematic analysis of patterns linked with transitional and temporal
elements of landscape systems (Fig. 4.1).
Nature can be understood as a complex collection of environmental operations
that the city has to support or to substitute like the filtering of water and air, the
provision of ecological niches, the growing of food, and many more. By viewing
the city no longer as a collection of independent objects of attraction but rather as
a systemic network of relationships in modulation of microclimatic conditions, we
can start a conversation on new more holistic approaches to future challenges related
to urbanization.
This approach aims at improved resilience of our urban environments and the
activation of cities as active generators of a balanced habitat. The subsequently
discussed design method is showcased through a series of computationally driven

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 49
P. Fricker and T. Kotnik, Patterns of Interaction,
SpringerBriefs in Architectural Design and Technology,
https://doi.org/10.1007/978-981-19-9083-0_4
50 4 Computing Land-Scapes

Fig. 4.1 Topological transformation of dynamic relationships: the system diagram is showcasing
the interaction of the existing field variables: landscape, environmental, and urban patterns within
a relational network system. Patterns and processes are brought in direct connection using dynamic
data interaction, which acts as a basis to define new design structures for the site supported by
information and formation (Image: Jenna Ahonen, Tina Cerpnjak and Feng Ye)

design speculations. The projects display speculations to a range of aspects, such as


climate change, environmental pollution, the loss of biodiversity and species, and
the exhaustion of natural resources in close cooperation with our environments.
These speculations are based upon a critical re-evaluation of predominant design
strategies from the digital realm, supported by an integration of ready-made algo-
rithmic tools and extending towards knowledge domains like thermodynamics.
Within the field of contemporary systems ecology, the potentials of thermodynamic
principles are further developed towards “the resilient self-organizing of ecosys-
tems” [4]. Such a discussion is based upon an extended awareness of the limitations
4.2 Digital Ecology Extended 51

caused by a static understanding of architecture and the narrow-minded belief in


performance-oriented solutions, supported by an increasing number of assessment
tools, such as “ecological footprints, carbon footprints, embodied energy, life cycle
assessment (LCA), cradle-to-cradle and ecosystem services review” [4].
Braham further discusses the discrepancy between rigid “forms of environmental
accounting” and meaningful design guidance, as part of an ongoing desire in defining
codes and standards to control dynamic global challenges [4]. These lines of crit-
ical thinking correspond to the introduced methodology and aim for a new open-
minded discussion on how singular buildings, parks, and green elements can trigger
a direction of connected productivity and system contributions. These dynamic living
networks shape new relationships, ultimately extending towards the formation of new
global systems, based on an understanding of the local environment and its inherent
patterns of interaction [15].
The integration of complex dynamic patterns as a starting point for computational
design thinking plays a central role in the understanding of the urban typology as
an articulated landscape. It enables an integrative systemic design approach across
scales and disciplines. Decoding site-specific systems, in relation to complex syner-
gies, fosters the closing of a gap between otherwise separated fields of knowledge.
The urban space should be created to expand the inherent logic and functionality
of nature and challenge the conventional understanding of densification strategies
enhanced by ecosystem services. Landscape and natural processes are understood as
a fundamental set of transitional concepts that operate on fields and boundary condi-
tions. Reading the urban fabric as dynamic network systems with different contex-
tual layers allows for an analysis of patterns linked with transitional and temporal
elements of landscape systems: interconnected networks reacting intrinsically to
dynamic change.

4.2 Digital Ecology Extended

As discussed in Chap. 3, the framework of patterns as a means for creative design,


supported by global and local, small- and big data packages, “must be rooted in a
discussion of system thinking and ecology” [23]. According to Pimm and Stuart, the
term ecology can be generally understood as “relationships between organisms and
their environment…. These interactions between individuals, between populations,
and between organisms and their environment form ecological systems, or ecosys-
tems” [27]. Since the 1960s, ecology is viewed as a science engaging in the general
understanding of environmental systems and the impact of these systems on change.
This new global ecological awareness was evoked by the pressure to develop new
methods to process the complex relations of nature-based processes, informed by the
availability of big spatial data. The famous image of Earthrise, taken by Astronaut
Bill Anders on a moonwalk in 1968, became a symbol of ecological consciousness
and an understanding of global context.
52 4 Computing Land-Scapes

If ecology and systems are common frameworks used to describe the constellations of rela-
tionships that we see in the world — the ‘what’ of the world — then patterns are the ‘how,’
or the means by which we come to know, understand, or express these relationships [23].

A case in point is the recent return to system thinking as applied to current


discourse on climate change and the broad field of sustainability in architecture
and landscape architecture. The power of system thinking methods and principles to
simulate complex environmental, societal, and political relationships found its first
global echo in the book The Limits to Growth [18]. Since then, a strong conscious-
ness of viewing the principles of ecosystems as potentials, where digitalization
becomes an integrative element has become a vital part in design and well accepted
in society [10]. The extended field of architecture, urban ecology, and many other
systems can be understood as patterns in a larger framework of system thinking
(Fig. 4.2). In order to reconnect isolated digital methods and be able to view them
across scales of action, designers are challenged to connect function and form with
respect to morphology, information, and communication, by integrating computa-
tional thinking to the large-scale observation of individual components in design
[23].
The importance of formulating strategies to confront the complex challenges of
climate change [21] has led to a second wave of ecological consciousness and a
subsequent renaissance of early computational theories of the 1960s and 1970s. In

Fig. 4.2 From separation to reformation: the diagrams display a series of temporal formations
of new relationships among design agents. Their dynamic behavior is providing a wide range of
adaptation strategies to be understood as an extension of nature in design (Image: Jenna Ahonen,
Tina Cerpnjak and Feng Ye)
4.3 Operative Extension of Nature 53

this chapter, the term digital ecology is viewed from a general, interdisciplinary point
of view as a new science, closely connected to biological and computational science
[9]. Due to the recurring notion of patterns, and its potential to inform applied ecology,
there has been a rich discussion regarding its application fields over the last 10 years.
In his article “The problem of pattern and scale in ecology: what have we learned
in 20 years?”, Jerome Chave points out the importance of extending the discussion
beyond energy and matter towards relevant challenges of scale and process:
The maintenance of ecosystem functions depends on shifts in species assemblages and on
cellular metabolism, not only on flows of energy and matter. These findings have far reaching
implications for our understanding of how ecosystem function and biodiversity will withstand
(or not) environmental changes in the 21st century (2013).

According to Chave, the discourse on patterns and scale has led to an important
enrichment within the field of ecology with respect to its history, which points out
the importance of understanding the intrinsic nature of systems and patterns with
respect to scale. Patterns and processes might react differently according to spatial
and temporal contexts, which is a very important element to be considered while
translating the behavior of specific patterns into an abstract computational design
description. In terms of scale, Chave discusses the technological advancement of
our time, like LiDAR and sensor technology, to capture global patterns and systems
and their relevance in distinguishing behavior and “spatial patterns of diversity”
from “global diversity patterns”. This excursion into the specific research domain
of digital ecology in relation to dynamic patterns is of huge importance for the
accuracy of developed computational models and often neglected due to a “language
boundary” between the disciplines. In order to understand the “how”, “patterns are
used as vehicles to understand, describe, and convey environmental processes” [7].
It is necessary to understand that the framework of dynamic patterns can only be
applied in terms of relationships, and this is the key for its power to be integrated
into architecture and landscape architecture across scales. The logic of a pattern
structure is not centered to any one object but branches out in a certain direction in
a characteristic manner. Due to the inherent and intrinsic character of patterns, the
logic can be applied to a diverse area of “natural and artificial systems” [7].

4.3 Operative Extension of Nature

In order to navigate the complex challenges facing the field of topological design
thinking, the introduced methods consist of an elaborated theoretical foundation,
which supports the development of strategies and tools within a flexible intellec-
tual framework. Unveiling the potential of computational design thinking, with
focus on dynamic patterns, offers the possibility to articulate design concepts based
on thinking in complex adaptive systems as opposed to predefined linear digital
workflows. These topological explorations are in dialogue with the early research
conducted at HfG in Ulm (see Sect. 3.2) and aim for a speculative exploration of a
54 4 Computing Land-Scapes

new topological design trajectory, driven by a computational mindset and comparable


to the radical movement fostered at the HfG in the 1950s:
Computing without computers allowed for the conceptualization of the computer, whereas
computing with computers transposed the traditional modes of operation and thinking into
a new medium [24].

In nature, complex processes and patterns of interaction are common. They enable
the configuring of heterogeneous physical elements into a coherent, homogenous
system and enable the emergence of new qualities by recombining existing elements
[23]. Similarly, exploring a new computational design strategy for integrating large
and small datasets into the design workflow harbors huge potential for entering a
computational design thinking methodology. As argued in Sect. 2.5, the field of
landscape architecture largely still operates in the digital era, focusing on a layer-
based approach, whereas the broad integration of information technologies lead to
a convergence of “diverse disciplines…using the same computational platform” [1].
The transgression of expertise into the territory of other disciplines questions the
nature of established boundaries. The breakdown of borders is enabled by informa-
tion technologies that operate with the same language across all disciplines, which
facilitates a transdisciplinary exchange of knowledge. We are now able to observe an
accelerated convergence of knowledge that results from facilitated communication
through a common intellectual platform (Fig. 4.3). The ability to work within other
disciplines through a common collaborative model encourages architects to increas-
ingly look beyond the discipline in order to investigate questions of morphology,
materiality, performance, sociability, and physiology [1].
Considering the importance of rethinking current linear design practices, the
conducted design speculations, creatively build up on the broad field of system
thinking and dynamic patterns in order to define a link between research and design
in the age of computation. According to Ahrens and Sprecher, “computational
systems promote the simulation of non-visual complex relationships, enabling a
synergy between the technology and the architect to generate the image” [1]. For
the presented discussion, the term “image” needs to be translated by “process—or
complex phenomena” and is formulated by Hedfors as follows:
The premise that transdisciplinary investigation of the abstract organization of phenomena,
independent of their substance, class, or spatial organization, reveals principles common to
all complex phenomena and provides a basis for models to describe and manage them [22].

According to Hedfors, “an important consequence of systems thinking has been


the shift from a purely quantitative view of the individual parts of nature to a more
qualitative understanding of nature as a whole and interactive system: a shift from
the singular focus on substance to a more balanced assessment of both the form and
the substance of interrelated phenomena” [22]. This view opens up to completely
new methods in order to reimagine relationships of sub-systems and their influential
environments to be applied as a guiding principle for computational design thinking,
in particular, for designing with an extended and operative understanding of nature.
4.3 Operative Extension of Nature 55

Fig. 4.3 An iterative design strategy stemming from a procedural simulation process towards a
process-oriented design method. Design is understood as an evolving and adaptive strategy, allowing
for interaction with known and unknown flow parameters in constant exchange with the time-based
approach of nature [6]

As discussed previously, systems thinking is inherently connected to patterns. For


the field of computational design thinking across scales, patterns articulate a visual
entry point for dynamic design solutions.
Observing phenomena within nature in terms of their inner logic reveals a multi-layered and
complex structure of their patterns. This thinking allows for connecting this inherent logic
to theoretical principles of computer science and to create a superstructure of computational
design methods [25]. James Corner similarly describes patterns as “relational frameworks
that simultaneously describe and project; they reveal structures, processes and relationships,
as well as structure physical frameworks that give shape and form to our world” [23]. The
logic of these connections and networks can be shown through patterns of behavior, which
manifest themselves in dynamic, active, binding, connecting, and distributing attributes.
Corner refers to the importance to relate this theoretical framework on the dynamic processes
inherent to landscape architecture, in order to “form new patterns and forms that structure
new ecologies, new program, and new modes of reception” [8, 13].

M’Closkey and VanDerSys show the potential of generative patterns for landscape
architecture and urban design to analyze structures and gain a fresh understanding
for relationships and the creation of form. The current development with respect
to the field of landscape architecture can be described from its integration in anal-
ysis to the connection of natural systems and their potential integration in design.
The behavior of natural systems and their dynamic nature is a key component. The
inherent dynamic and emergent character of nature can be described through the
phenomena of “self-organization and emergence” [23]. This extended definition of
56 4 Computing Land-Scapes

emergence is understood within the realm of landscape architecture, “to material and
cultural transformations that are presumed to occur after design implementation”
(Fig. 4.4) [23].
Likewise, James Corner states that “a truly ecological landscape architecture
might be less about the construction of finished and complete works, and more
about the design of ‘processes,’ ‘strategies,’ ‘agencies,’ and ‘scaffoldings’—catalytic
frameworks that might enable a diversity of relationships to create, emerge, network,
interconnect, and differentiate” [23].
Computational design is primarily about the detection of patterns as “the ‘how’ or the means
by which we come to know, understand, or express these relationships” [23]. Such thinking in

Fig. 4.4 Abstraction of river dynamics into interactive particles: the dynamics of water flow, annual
cycles, sedimentation behavior, and urban growth is decoded into a computational system, which
is visualized through patterns of interaction. Focus is set on the definition of temporal patterns and
opportunities for interaction in order to detect potentials for symbiosis between the urban and the
landscape (Image: Janne Keskinen and Yinan Xiao)
4.4 On the Notion of Flows 57

patterns places computational design in conceptual proximity to mathematics: mathematics


is the science of patterns, whereby fundamental patterns result from the formalization of
human perception. Accordingly, from a theoretical point of view, computational design is
grounded in structuralism. Not in an anthropological structuralism based on linguistic studies
of Ferdinand de Saussure but rather in a structuralism of the natural and technical sciences
is grounded in Norbert Wiener’s studies on cybernetics [31] and the work of Ludwig von
Bertalanffy on general system theory [30, 15].

Our world can be described through patterns. This can be viewed on a purely
formal visual level as well on a much more complex one with respect to underlying
systems and their relationships. “Patterns in designed landscapes are often under-
stood as implying the imposition of order, reflecting human dominance over the
complexities and flux of nature. At times, they are equated with static surfaces, such
as parterres and paving patterns; in other instances they are associated with the repet-
itive configurations of urban or agricultural land use. Yet the importance of patterns
goes well beyond such readily recognizable formal attributes as simple surfaces or
uniform geometries” [23].
With respect to data and computational processes, the integration of patterns
supports an understanding of processes in nature since they represent the interplay
of material and forces across time and scale.
Using patterns with respect to data visualization, we gain an automatic visualization of
processes, currents, and entanglements of interrelated systems. The implementation – or
decoding – of these visually represented rules, can be used to create new dependencies,
as described by [3], to be used directly as design instruments. Computation supports an
understanding of complex natural behavior, which in return informs the designer of a new
understanding of data. Belesky describes these informed methods as a possibility “to analyze
and generate geometric information as a sequence of logic procedures, and to begin to close
the gap between what constitutes a ‘tool’ and a ‘technique’” [12, 29].

4.4 On the Notion of Flows

In order to define drivers for change, computational design principles need to have
an appropriate methodology in relation to theory and application areas, which goes
beyond the definition of workflows.
The relevance of Topological Design Thinking for the extended field of architec-
ture offers an avenue to reconnect the fragmented digital design approach by viewing
computational design thinking as an intellectual process. In order to take advantage
of computing power, the process of interacting with complex phenomena articulates
design through a purposefully designed code and not through the integration of off-
the-shelf digital tools. Understanding coding as a creative translation or mediator
between data (or information), scientific methods, processes, and the creative gener-
ation of design allow for the field of design to go beyond the discussion of a static
form or image. Programming directly connects to the discussions of Wiener in rela-
tion to information and formation [32]. Gilbert Simondon reflects on this connection
to cybernetics and the question on “how to connect information to formation, and
thus link information to form” as follows:
58 4 Computing Land-Scapes

[T]he meaning of information, therefore, has switched toward information, in the realm of
form-giving. Accordingly, a system with an elevated quantity of information is an ordered
system, presenting significant, meaningful forms, as opposed to a state of undifferentiated
disorder [28].

Relating to this line of thought, Ahrens and Sprecher highlight the unique current
possibility for our discipline, as information technologies “has been carried back into
the architectural profession” [1]. This observation is partly true but especially within
design, we are partly still concentrating on the codification of an image, rather than
integrating non-linear design principles.
As the field of architecture is challenged to increasingly interact with dynamic
and temporal environmental components, the reflection on the difference between
a linear and a circular process in relation to computational design thinking is of
fundamental importance.
One important motivation for the critical re-evaluation of current mainstream
sustainable design methods is the viewing of building as an extension of nature,
responsible and capable of not only consuming energy and matter but also to produce
and reproduce itself. In his book A city is not a tree, Alexander critically reflects on
the difference between “natural cities” and “artificial cities” from a computational
point of view [2]. Alexander’s title has to be understood as a criticism on the plan-
ning system of the time, as he highlighted the danger to structure a problem using
a “tree-structure” [16]. Alexander’s theoretical finding highlights the importance to
extend the understanding of urban structures beyond a set of singular elements, but
rather understood as a complex system. Current generative design approaches display
avenues towards the formulation of answers to the critical findings of Alexander who
addresses the “rise of the need for an advanced model, driven by the rules of math-
ematics and cybernetics, based on interdependence and on the value of feedback”
[16]. The potential of generative design methods can be further extended towards a
new role of the architect, understood as an informed curator and “caretaker” versus
top-down decision-maker (Fig. 4.5).
Extending this discussion towards the initially mentioned formulation of orga-
nized matter (see Sect. 3.4), viewed as thermodynamic reading of architecture as
organized energy flows [11], the systemic understanding of the basic physical prin-
ciples of the elements, sun, wind, and water is further researched about regarding
on their potential to act as a basis for speculative computational design articulations
[15]. The discussion on processes by Fernández-Galiano describes the relationship
towards energy from the maintenance and construction point of view and critically
outlines “the scandalous absence of energy considerations in architectural analysis
and criticism” [11].
The irruption of energy in the universe of architecture smashes its crystalline images, shakes
its mute silhouette, and gives it a definitive place in the field of processes and life. Architecture
can then be thought of as a transformation of the material environment by changing living
beings, an artifact continuously altered by use and circumstance, in constant degradation
and repair before the aggression of time, permanently perishing and renewing itself [11].

Our environment is confronted with the need to transform, adapt, and react to a
magnitude of unknown challenges, which require radical methods for environmental
4.4 On the Notion of Flows 59

Fig. 4.5 For a housing project at Hanasaari, Helsinki, an optimization procedure is used for the
study of generative design alternatives. In the first study (upper two rows), a pixelated volume is
generated with the aim to maximize the solar radiation amount both for diminishing heating loads
and provide direct sunlight while retained increased visibility for residents. In the second study
(lower two rows), the main goal is to create a series of shared large communal semi-private terraces
that can involve residents to activities and self-organized space distribution. At the same time, the
buildings should be kept compact by increasing usable area in combination with minimized surface
area, thus reducing heat loss (Image: Dan Palarie)
60 4 Computing Land-Scapes

interventions understanding the potential of “free energy”. The schematic arrange-


ment of matter, and the flow of energy, which adapts to the environmental condition
can be exemplarily researched, analyzing the elements, sun, wind, and water with
respect to their specific physical properties and behavior. The subsequently intro-
duced computational design strategy engages with these behavior patterns giving rise
to unconventional open systems design thinking, based upon energy flow feeding the
process [11].
The project Constructed Ground—Hawker Centre Park (Fig. 4.6) promotes a
promenade of transitional spaces flowing through an articulated topography as a
vector stream, favoring the appearance of programmatic conflicts at various levels in
order to activate the necessary convergence among activities in a topological design
manner. With that purpose in mind, a number of spatial operations, such as people,
wind, and water flow, were identified and thoroughly explored as the base for the
development of the articulated landscape.
By means of an apparently mechanical process of disciplined play, sequences of
rich and varied architectural experiences are meticulously created. Using variations
in proportion and spatial strategies of simple parametric pattern compression and
expansion, collective spaces are articulated as intrinsic parts of the Hawker center. In
addition, cross ventilation and self-shading principles are implemented simultane-
ously as a part of the space-making strategy—strategies of importance in the tropical
climate.
Situated near the equator, Singapore has a tropical climate with uniformly
high temperatures and humidity all year round. As such, a large portion of elec-
tricity consumption is typically used for space-cooling. A typical household energy
consumption profile shows that air-conditioner’s account for the largest proportion of
total electricity consumption. With increased urban growth, electrical consumption
will only continue to rise. As an effort to reduce air-condition usage, the project seeks
to utilize passive cooling strategies in building design to create a comfortable natural
environment for its users (Fig. 4.7). Grounded in the analysis of local conditions on
site, the focus of the design was on harnessing air flow. The goal was to maximize
wind entering into the building from the macroscale down to the micro-scale. Apart
from wind, secondary strategies of self-shading and controlled circulation by means
of ground articulation were included in the holistic cooling strategy.
The articulated ground condition is the fundamental point of departure for the
topological design concept of the project Intercellular–Computational Territories
(Fig. 4.8). The terrain and the built elements are part of a computational design
thinking stimulated by the flow of wind, water, and people, allowing for a new urban
typology seen as urban field conditions, to be discussed. The artificial peninsula
Hernesaari, a former industrial area in Helsinki (Finland), serves as a prominent
testing ground currently going through a redevelopment by the municipality. The
project translates behavioral patterns of water and wind into a responsive open system
of organized matter. This process-oriented design method simulates a new interac-
tion with site-specific parameters, like the change of seawater levels, on-site water
purification, and changing wind conditions, by mixing form and functions into a new
spatial configuration. “The zoning of the site is controlled by a cellular system, using
4.4 On the Notion of Flows 61

Fig. 4.6 Constructed ground—Hawker Centre Park (Image: Pheeraphat Ratchakitprakarn and Ho
Yu De Samuel)

the history, the present and a speculative future situation as visionary input parameter.
Applying this method, a highly flexible land system is generated to realize multipur-
pose functions, merging environmental aspects with a future-oriented urban strategy.
The topography of the articulated ground condition is automatically generated and
influenced by environmental factors like rain, wind and sun direction. Through this
method the runoff water is purified in responsive water catchment areas, which as
well interact as a mediator for different seawater levels. A systemic interplay of
people, natural elements, and buildings is generated as a logic consequence of the
computational model allowing for change and adaption” [12, 13].
The aforementioned design speculations support a discussion aiming for bridging
between the field of “energy that accumulates information” [11] and topological
design methods. The discussed architectural articulations extend beyond the umbrella
terminology of sustainability and connect to the conversation on open thermodynamic
62 4 Computing Land-Scapes

Fig. 4.7 Landscapes of passive cooling—Dense Vertical Living (Image: Eva Yang, Grace Teo and
Megan Chor)

systems which “clearly spell out the relation between form, matter, and energy: the
capacity of matter to accumulate energy as in-formation, and the need for matter
to receive energy to maintain its form [11]. Matter, hence, needs energy in order to
maintain its form, and form, in turn, can be thought of as a wealth of stored energy”
[11]. This way of thinking relates and supports the generation of spatial organiza-
tions across scales and sub-systems, viewed as a process and closely linked to the
discussion on living organisms, as introduced through the comparison of natural and
manmade systems by Frei Otto in Sect. 3.4. This notion has been recently further
discussed from the angle of eco-thermodynamics and stated by Kiel Moe as: “Situ-
ating architecture in these hierarchies poses new questions about energy for designers
that far exceed the common preoccupations with conservation, efficiency, and opti-
mization that dominate the discourse on architecture and energy” [20]. This abstract
conversation is supported by the tangible field of dynamic patterns, as a medium
4.4 On the Notion of Flows 63

Fig. 4.8 Intercellular–Computational Territories (Image: Jiaqi Wang, Xin Ding and Pirita
Meskanen)

to translate energy accumulated as matter in a closed system, towards a process of


constant exchange with the environment, showcased through the relationships of
flows across scales. This results in a computational design teaching methodology,
based upon pattern generation, pattern transformation, and speculative interpretation:
Pattern Generation: Taking the structuralist perspective as a framework, the projects research
into the local parameters of the site (urban growth, flow, sedimentation, water dynamics,
human factors) in order to formulate an underlying systematic approach for translating these
findings into abstract patterns. Special focus is set on the generation of patterns for process
and performance [5].
Pattern Transformation: Within the second part, abstract design patterns will be trans-
formed, juxtaposing the theoretical inputs from landscape architecture and urbanism and
computational input of the course. In designing an artificial manmade urban landscape
with natural elements, we will examine and manipulate topological mechanisms [17]. In
understanding and adjusting these mechanisms, we will provoke physical reactions that will
64 4 Computing Land-Scapes

structure the site and provide the basis for a new urban pattern [19]. Design scenarios will
be organized along both physical and temporal lines. Strategic design and system thinking
must enable variability in form, shape, and scale.
Speculative Interpretation: In the final step, students are asked to formulate their own
speculative design hypotheses by formulating a future-oriented approach towards the chal-
lenges of the site based on their pattern exploration. The aim of the speculative design
hypotheses is to structure and prepare the site for future developments through an integrated
design approach [14]. The speculative design hypotheses will support sustainable urban
growth and the intensification of urban areas as well as encourage a dynamic interaction
with underlying potentials or parameters of the site [15].

4.5 Umweltecture—Sustainable Visions Between


Architecture and Landscape

This chapter initiates a comprehensive and future-oriented discussion for possible


avenues towards retooling current digital design towards a computational design
thinking methodology, supported by topological design thinking, and showcased
through a series of design speculations. The speculations do not aim to focus on
specific tools or workflows, but rather they deal with a holistic framing of the
problem (discussed in Chaps. 2 and 3), underscoring the importance of getting back in
command. If, as suggested by Sarah Williams, data can be regarded as the new infras-
tructure, the field of architecture is able to envision new ways of how to creatively
interact with new methods from expert systems like Artificial Intelligence (AI) and
Machine Learning (ML), robotics and mixed reality, in order to finally define adap-
tive solutions for the grand environmental, societal, and geopolitical challenges we
face today [33]. As stated by Picon already in 2015, “the entire city could be consid-
ered intelligent in a new way, founded on the interaction and composition of the
perceptions and deliberations of multiple entities, human, non-human, and often a
mix of the two” [26].
The emergence of Big Data and implicit discussions of their relevance in the
area of smart cities marked a turning point in architecture as well as landscape
architecture. Strongly influenced by themes from AI and the fusion of virtual and
real environments, we currently find ourselves in a situation where the possibilities
of application become almost infinite; where not only the upscaling of experiments
stand at the forefront, but a deep dispute over future-oriented research directions.
Digital material gradually lost its relation to real scale, input and output was discussed
without relation to a specific place. An essential aspect was also a distancing from
theoretical discourse, which is why digital design was pushed more and more to
the periphery of teaching and regarded merely as a tool. The current increase of
innovation und experimentation in the area of data interaction in architecture is due
to the complexity of current environmental challenges and could, similar to the era
of parametricism in architecture, lead to positive new directions in design (Fig. 4.9).
4.5 Umweltecture—Sustainable Visions Between Architecture and Landscape 65

Fig. 4.9 Urban Rhizome: extension of computational design methods driven by “re-framing of
architecture as a practice dealing with the design of discrete artifacts, to a process of shaping and
diverting energy flows. The theory of thermodynamic acts as a base for understanding such processes
of degradation and confirmation, making them available as operative tools in the design process”
[6]

The introduced design speculations are taking their point of departure from the
aforementioned observations, focusing on understanding the power of data inter-
action in relation to the overarching discussion on flow dynamics (Fig. 4.8). The
developed theoretical framework serves as a basis for all of the introduced design
methods. The current design challenges are part of a highly dynamic tension field,
working with complex processes and dynamic interrelationships between different
systems, which are often difficult to predict. In order to face the future challenges of
our profession, design thinking has to incorporate multidisciplinary aspects as well
as focus on the element of dynamic simulation and automated iterations. Compu-
tational design thinking is not a set of digital tools but an approach that empowers
designers to think and design in larger systems by identifying the inherent logic
of local patterns guided by aesthetic principles for collaborative futures beyond the
Anthropocene.
Architecture will move towards Umweltecture (referring to Uexküll’s concept of
Umwelt), or “the cultivation of an environmental culture”, and allow us to engage with
surrounding urban and green systems and transform these concepts into a new urban
typology grounded in social and environmental sustainability. This convergence of
architecture and landscape will establish a reflection on a “new nature”: a mix of
technology and the natural, emphasizing nature’s malleability and in terms of human
intervention, inherent artificiality.
66 4 Computing Land-Scapes

References

1. Ahrens C, Sprecher A (2019) Instabilities and potentialities: notes on the nature of knowledge
in digital architecture. Routledge, Taylor & Francis Group, New York
2. Alexander C (1965) A city is not a tree. Archit Forum 122(1):58–62
3. Alexander C (1967) Notes on the synthesis of form. Harvard University Press, Cambridge,
Massachusetts
4. Braham WW (2015) Architecture and systems ecology: thermodynamic principles of environ-
mental building design, in three parts. Routledge
5. Cantrell BE, Holzman J (2015) Responsive landscapes: strategies for responsive technologies
in landscape architecture. Routledge
6. Cerpnjak T (2021) Formal diffusions. Agent-based-systems in the context of architectural space
making
7. Chave J (2013) The problem of pattern and scale in ecology: what have we learned in 20 years?
Ecol Lett 16:4–16. https://doi.org/10.1111/ele.12048
8. Corner J (1997) Chapter 3, Ecology and landscape as agents of creativity. In: Thompson GF,
Steiner FR (eds) Ecological design and planning. Wiley, New York
9. Dodds WK (2009) Laws, theories, and patterns in ecology. University of California Press,
Berkerley, UNITED STATES
10. Egan M (2007) Barry Commoner and the science of survival: the remaking of American
environmentalism. MIT, Cambridge, Mass
11. Fernández-Galiano L (2000) Fire and memory: on architecture and energy. MIT Press,
Cambridge, Massachusetts
12. Fricker P (2022) Computing with nature: digital design methodologies across scales. In: Mona-
cella R, Keane B (eds) Designing landscape architectural education: studio ecologies for unpre-
dictable futures, 1st edn. Routledge, Taylor & Francis Group, Milton Park, Abingdon, Oxon;
New York, NY, p 438
13. Fricker P (2022) Augmented co-design methods for climate smart environments: a critical
discourse and historical reflection. In: Ugliotti FM, Osello A (eds) Advances in human and
social aspects of technology. IGI Global, pp 156–183
14. Fricker P (2016) Extending the limits: using big data as integrated design tool within the field
of large-scape landscape architecture. In: Education for research, research for creativity—
architecture for the society of knowledge. Warszawa, pp 30–36
15. Fricker P, Kotnik T, Piskorec L (2019) Structuralism: patterns of interaction computational
design thinking across scales. Wichmann Verlag, DE
16. High Dutton Associés (2013) Christopher Alexander: “A city is not a tree.” In: Complexitys.
https://complexitys.com/2013/05/16/christopher-alexander-a-city-is-not-a-tree/. Accessed 20
May 2022
17. Leach N, Yuan PF (eds) (2018) Computational design. Tongji University Press Co. Ltd,
Shanghai
18. Meadows DH et al (1972) The limits to growth: a report for the Club of Rome’s project on the
predicament of mankind. Earth Island, London
19. Menges A (2011) Computational design thinking. Wiley, Chichester
20. Moe K (2013) Convergence: an architectural agenda for energy. Routledge, New York
21. Moss RH, Edmonds JA, Hibbard KA (2010) The next generation of scenarios for climate
change research and assessment. Nature 463:747–756
22. Murphy MD, Hedfors P (2011) Systems theory in landscape architecture. In: Conference: urban
nature: Council of Educators in Landsacpe Architecture CELAAt, Los Angeles, California, pp
1–13
23. M’Closkey K, VanDerSys K (2017) Dynamic patterns: visualizing landscapes in a digital age,
1st edn. Routledge
24. Neves IC, Rocha J, Duarte JP (2013) The legacy of the Hochschule für Gestaltugn of Ulm for
computational design research in architecture. In: Stouffs R, Janssen P, Roudavski B, Tunçer B
(eds) Proceedings of the 18th international conference on computer-aided architectural design
research in Asia. NUS Singapore, pp 292–302
References 67

25. Picon A (2010) Digital culture in architecture: an introduction for the design professions.
Birkhäuser, Basel
26. Picon A (2015) Smart cities: a spatialised intelligence. Wiley
27. Pimm S, Smith RL (2019) Ecology. In: Encyclopædia Britannica. https://www.britannica.com/
science/ecology. Accessed 23 Apr 2020
28. Simondon G (2019) Foreword: Gilbort Simondon’s key points. In: Ahrens C, Sprecher A
(eds) Instabilities and potentialities: notes on the nature of knowledge in digital architecture.
Routledge, Taylor & Francis Group, New York, pp IIX–XXII
29. Tara A, Belesky P, Ninsalam Y (2019) Towards managing visual impacts on public spaces:
a quantitative approach to studying visual complexity and enclosure using visual bowl and
fractal dimension. Wichmann Verlag, DE
30. von Bertalanffy L (1969) General system theory foundations, development, applications.
GBraziller, New York
31. Wiener N (1948) Cybernetics, or control and communication in the animal and the machine,
Print 6. Hermann, Paris
32. Wiener N (1989) The human use of human beings cybernetics and society. Free Association,
London
33. Williams S (2020) Data action: using data for public good. The MIT Press, Cambridge,
Massachusetts
Epilogue

In the past three chapters, we have argued our idea of a design approach in architec-
ture, urban design, and landscape architecture based on the exploration of patterns
of organization that expand the order of natural systems into the built environment.
In Chap. 2, we discussed the transformation of the notion of context into a data-rich
ground and the resulting convergence of design methods and strategies on various
scales of architectural thinking. These layers of data are related by a network of
relationships that can be made operative for design by a cascading sequence of topo-
logical schema, or patterns of interaction, discussed in Chap. 3. In the subsequent
chapter, we illustrated the impact of this design approach on all scales of architecture
through a number of design explorations.
Like all research, the ideas we have presented are not originalis, not first in time.
Rather they build on a multitude of ideas from other thinkers, especially from around
the 1960s. This time period saw the beginning of an environmental consciousness and
the building up of a movement to address ecological problems caused by pollution,
suburbanization, and industrial agriculture. In the United States, this movement led
to landmark legislation, including the National Environmental Policy Act (1970), the
Clean Water Act (1972), the Endangered Species Act (1973), as well as the creation
of the Environmental Protection Agency (1970), the very first institution of its kind.
But this period of time is also of great interest for the growth of architectural research
that represented a shift away from the traditional conception of the architect as an
artist-craftsman and the first formulation of science-based design methodology.
Unfortunately, most of these promising beginnings slowly passed into oblivion
over time. Over half a century later, ecological problems have scaled up into a global
climate crisis and the discussion of an architectural epistemology has resurfaced
with digitalization and its widespread social consequences. With the urgency of this
research in mind, central ideas the authors discussed with two experts in landscape
and environmental design: Emanuele Naboni and Christophe Girot.
Emanuele Naboni is Professor at the University of Parma, affiliated Professor at
the Royal Danish Academy in Copenhagen and Principal of a consulting company

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 69
P. Fricker and T. Kotnik, Patterns of Interaction,
SpringerBriefs in Architectural Design and Technology,
https://doi.org/10.1007/978-981-19-9083-0
70 Epilogue

on sustainable design. He was a Researcher at the Lawrence Berkeley National


Laboratory (LBNL) and Performance Designer at Skidmore, Owings and Merrill
(SOM) in San Francisco. Naboni was visiting Professor at the Singapore-ETH Center
Future Cities Laboratory, an invited Professor at Ecole Polytechnique Fédérale de
Lausanne (EPFL), Architectural Association (AA), and UC Berkeley. His expertise
is in sustainable design solutions, strategies, methods and digital simulation seeking
adaptation to climate change, ecosystems quality, sustainable performance of urban
environments/buildings, and human health.
Christophe Girot is Professor and Chair of Landscape Architecture at the Depart-
ment of Architecture of the ETH in Zürich since 2001. His research is focused on
large-scale landscape design and modeling methods with particular attention to the
topology of nature in and around cities. The LVML (landscape visualization and
modeling lab) of ETH, funded by the Swiss National Science Foundation and shared
by the Department of Architecture and the Department of Civil Engineering and
Geomatics, has enabled significant advances in applied landscape design and point
cloud modeling. Ongoing research has yielded groundbreaking results in point cloud
design, modeling, and acoustic sensing. Girot has a practice in Zurich with projects
in Europe and in Asia. The Sigirino Mound for the Alp Transit Company in Ticino
as well as the Brissago Garden project with SAM Architects test current limits of
topological design and modeling in challenging alpine contexts. His many important
publications include The Course of Landscape Architecture, published by Thames
and Hudson in 2016.

A.1 Interview with Emanuele Naboni

The notion of sustainability has gained increasing importance in architecture and


urban design over the past few years. There is an awareness that every impact of an
intervention needs to be reviewed with respect to its impact on the environment, and
that this should be a crucial part of every design consideration from now on. Inter-
estingly, in your recent book Regenerative Design in Digital Practice, you criticize
the widespread focus on sustainability, or to be more precise, the current common
understanding of sustainability as being inadequate. I would be very interested in
learning why you think this is the case?
In conventional sustainable design and sustainability research, most discussions
are driven by checklists in search of a label that authoritatively certifies the acquisition
of a certain kind of “status”; solutions are applied like a recipe that lists the ingredi-
ents without any further guidance. On the one side, design is becoming fragmented
when subjected to off-the-shelf technology blindly supporting generally sustain-
able targets: buildings and cities often end up looking like a collection of technical
gadgets. On the other side, the targets are seldom ambitious and propose reduc-
tionist approaches. In contrast, I propose three exemplary projects related to cooling
cities, building envelope insulation, and daylighting. Specific research is focused on
cooling down cities by the means of cool roofs and cool pavements, both featuring
Epilogue 71

low albedo. In this case, the focus is exclusively on horizontal surfaces, completely
neglecting the impact of facades, which are the primary surfaces absorbing the
highest rate of longwave thermal radiation and reflecting solar shortwave radiation
towards walking pedestrians. This is a gross error that impacts pedestrian comfort
significantly, as perceived temperatures can vary up to 12° depending on envelope
design. If we move to the scale of the building, the impact of climate change is
often negligible, and we are still talking to a large extent about insulation. All the
European directives and the certification systems and energy modeling-based ratings
are still primarily referring to a winter scenario, and meteorological data commonly
used in energy analyses is often not up to date. This results in buildings that are
neither able to cope with the future, nor even today’s climate. This situation leads to
the current trend of building extremely well-insulated buildings, which have much
higher cooling demands in today’s warmer climate. The overuse of insulation in
passive building standards, for instance, causes a greater overheating of the building
and increasing hours of discomfort. A study we conducted with Lund University
showed that building retrofits in Germany and Denmark based on supplementing
insulation are dramatically increasing indoor discomfort hours. Lastly, when we talk
about daylighting, norms and rating systems tend to focus on narrow discussions
about quantifying daylighting with metrics such as useful daylighting illuminance
and similar measures. But occupants are not seeking merely certain quantities and
intensities of light. The use of these guidelines is “flattening” the light quality in
contemporary architecture around the world. This is a pity, as a much more extended
light vocabulary, including aspects of contrast and temporal and spatial variability
under daylight conditions, stimulates occupants’ engagement with their surround-
ings. Although subjective, the perceptual performance of space should rank above
mere quantitative measures. I studied modernist and contemporary Scandinavian
architecture in depth: the designs of Alvar Aalto and Ralph Erskine would never
fit any contemporary metric, yet their architecture are masterpieces of light: the
subtle differentiation of light directionality, or the reflection of light as it comes
from the outside to the inside, in their designs is outstanding. Today, overruling
light with scientifically determined norms is reducing the daylighting discourse to a
mere number game. To sum up, the type of scientific and applied targets commonly
discussed in sustainability are limited in scope, and there is a big risk of producing
average solutions partially able to fulfill sustainability requirements but not able to
promote a high qualitative and artistically rich set of solutions.
Looking at all these examples that you mentioned and discussed, it seems that
your critique focuses on a general lack of systemic thinking. Our current approach
to sustainability is still very much following a paradigm of design thinking that was
typical for the twentieth century; a functionalistic way of thinking where we isolate
a problem and then try to develop solutions for it. But this cause-and-effect scheme
falls short of any systemic understanding: the systemic is not part of the notion of
sustainability as it is used in design disciplines today. It seems that your critique
implies that sustainable design requires a paradigmatic shift of our understanding
of the manmade or how the built environment, in the larger sense, interacts with
the natural. In some way, sustainability should no longer focus on something like
72 Epilogue

limiting impacts, but rather more on maximizing the way something can enhance or
support the natural system. Is this a correct interpretation of your views?
Until now, sustainable building design in Europe and North America has been
driven by the concept of a hyper-insulated system with radiant surface systems for
heating and cooling. These systems aim at thermodynamically separating the indoors
from the outdoors. This combining of insulation with temperature regulating mech-
anisms did much more than thermodynamic machines to disconnect buildings from
their sites. These systems attempted to “Design with Climate”, to cite the title of one
of the most known books in the field. But design for climate change and ecological
uncertainty implies embracing an emerging worldview paradigm, where architects
make systemic concessions to the climate and where the outside, the larger climatic
and ecological context, is part of the design. There are already ideas of buildings
offering climate mitigation and enhancing biodiversity, generating clean energy, and
growing healthy food. Materials should be planned for infinite life cycles, build-
ings could ideally remediate air, water, and soil, while simultaneously contributing
to meaningful cultural and social experiences. Incorporating such thinking means
that designers must grapple with more information than ever before and, to add
complexity, they must operate without future climatic certainty. While climate-
change-related approaches are attracting growing interest, transitioning to an effec-
tive regenerative practice remains a challenge. Design for climate change implies
a much more systemic analysis of problems and the formulation of complex multi-
domain research. This implies that architects will have to blend scientific approaches
into a holistic design framework: a phenomenal task for architects. To prepare archi-
tects for this challenge, I have collaborated with David Garcia at the Master of
Architecture and Extreme Environments at the Royal Danish Academy in Copen-
hagen to develop scientific skills by promoting research methods like data collection
and mining, and data discussion as the main raw materials for design (Fig. A.1).
From your perspective there is no need for a revolution in design thinking in
the sense of a radical paradigm shift. You are rather optimistic that the continual
learning about our world will encourage the human-centric perspective to be more
inclusive. What I find interesting is that all of your answers reflect strong scientific
underpinnings. Your approach is very much anchored in a scientific worldview. For
you, science is one of the important drivers for the evolution of design thinking, and
the integration of more scientific knowledge is an important aspect of achieving more
sustainable design. In your practice, one of the crucial elements enabling such scien-
tific grounding is the use of simulation tools. Traditionally, however, simulation has
been primarily used as an analytical tool. The challenge: How does this simulation
get operative in the sense of initiating positive change? How can a simulation be a
tool that solves, or at least supports the kinds of issues that you were touching on
and be forward-looking and not only analytical in the sense of looking at what we
have at the moment?
In practice, I often found simulation modeling very primitive: often we were forced
to deal with one single environmental issue at a time. Most tools are concerned with
the reduction of operational and embodied energy consumption and emissions; the
optimization of indoor thermal comfort and visual comfort, or the modeling of air
Epilogue 73

Fig. A.1 One of the first large-scale models allowing a high-resolution understanding of tempera-
ture gradients between indoors and outdoors. In the model of parts of Gothenburg, Sweden, indoor
and outdoor are thermally linked. For instance, it is possible to link the indoor operation of buildings
to microclimate effects of a local forest, or it is possible to estimate the impact of temperate seawater
on the indoor temperature of adjoining buildings. Each tree and vegetation species is modeled, as
well as the type of ground and building properties. The model is calibrated with on-site recordings.
Throughout a stochastic set of equations, it is possible to predict variations in temperature due to
climate change in 2050 in the territory of Gothenburg. The chart functions as a base for future
planning, showing how local forests and wind channels need to be preserved and suggesting what
types of buildings should be adapted to new climatic trends. The model illustrates that most of the
buildings will be uninhabitable in summer times (Image: Emanuele Naboni, Antonello di Nunzio
and Thomas Amlov)

flows. I personally aim for regenerative and systemic design and have found it diffi-
cult to find simulation tools that are able to be customized in sophisticated ways. I
want to respond to a much broader set of science-based targets, such as those of local
ecosystems and human health, as well as tools that are able to handle interrelated
issues in a more sophisticated manner. We need to create new tools. A month ago, we
developed a temporary pavilion in a historical courtyard in Ascoli, Italy. Our work
started by identifying local degeneration processes. We attempted to understand local
climate, ecology, carbon cycles, and potential health issues of the inhabitants. We
performed a systematic overview of local data collected with Arduino basic sensors
and bolstered this information with qualitative surveys. Data related to local climate,
local water cycles, the behavior of other species, and natural patterns of vegetation
growth and human physiology helped us determine relationships, and we corre-
lated all phenomena to climate change information. We wrote custom components
to include multi-domain equations for ecosystems, buildings, and spaces as well as
inhabitants as a function of future temperature variations. The final design promotes
with one design move several benefits in biodiversity, CO2 sequestration, air purifi-
cation, circularity of materials, and thermal resilience to current and future climates.
Although it is a small and transient structure, we believe it sets an example of a multi-
domain and multi-simulation approach with extended means and a strategy that can
be used to design for climate change. The challenge is to replicate this process in
future and more prominent architectural interventions. We have now signed a contract
with the Trenitalia (the Italian national railway) to scale this approach to new and
74 Epilogue

climate change-adapted train stations, which will regenerate local ecological systems
and foster overall well-being.
Your discussion of the courtyard project illustrates the complexity not only of the
situation itself but also of the methodology. In practice, we often see an unshakeable
belief in quantification, a belief in the use of simulations but also in parametric tools in
general. The seemingly objective number supports a tendency to work out a solution
based on specific boundary conditions, and its validity is arbitrarily guaranteed
through the quantitative approach.
But if we come back to the systemic perspective, we know that in complex systems
often found in nature or in urban situations, minor changes in the parameters can
have quite a huge impact on the outcome or in the way the system starts to behave.
With this in mind, I wonder if we do not have to foster much more of a qualitative
perspective, instead of focusing on one specific calculated and thereby infallible
solution. Maybe we need to focus on larger sets of scenarios to consider a range of
behaviors. I think we need to expand our view of space and time when we analyze
in the manner you describe. In using computational tools, we should not aim at the
solution, but rather at a range of solutions or scenarios, or, in other words, instead
of looking for specific numbers, we are rather looking for patterns in the behavior,
in what manner things are changing. And these patterns could be design drivers to
help us understand how we as architects can influence that general behavior. What
we as architects should be aiming at is trying not to fall into this pitfall to achieve a
specific result, but rather encourage a specific behavior or pattern of behavior.
I believe we have to start using these computational tools in a more qualitative
scenario-based form of reading, instead of aiming for a specific solution for one
number set. The workload will increase enormously, but it seems to me that the only
way of really working with these tools is in this kind of sensible way. Would you
agree?
When we think about the design of our built environment, we can recall interven-
tions driven by set of simulation data leading to a specific solution and technology.
The broad outcomes and the nature of climate change breakdowns cannot be under-
stood solely through numbers (e.g., CO2 emission reduction, air temperature miti-
gation, etc.). Silo thinking and purely numerical approaches are dysfunctional in a
changing climate as different criteria are continually in tension. Data and simulations
may be integrated by a local reading of nature’s every aspect. Seasonal cycles keep
retracing the sites. Even when we try to govern these circular natural flows, they
find ways to manifest themselves. These transient but contextual relationships with
nature are a hugely important layer of understanding in reading a place. Such obser-
vations may be relevant inputs for informing simulation workflows, where design
is mediated by site features. The holistic understanding of local phenomena, forces,
or atmospheres can also qualitatively inform the simulation trajectory. Part of the
answer may lay in the definition of parametric design. In modern Latin, parametric
means “for the measure” from the Greek para, or “beside”, and metron, or “mea-
sure”. Design comes from the Italian verb disegnare from the sixteenth century,
which means both “to develop the senses” as well as “to contrive, plot, intend”. The
French took both meanings from the Italian, in different forms and passed them on
Epilogue 75

Fig. A.2 The Lavazza building facade is an example of an envelope where flows of radiation,
light, and views led to a vocabulary of heterogeneous facade elements. In the original consultancy,
our simulations showed that it was possible to mediate solar radiation in a differential way so
that we could change indoor thermal conditions and light for each of the rooms. Whereas norms,
certification systems and mechanical engineers’ proxy would recommend fixed indoor conditions,
this work was inspired by the most recent scientific research on the concept of thermal variation.
Thermal variability benefits health and well-being, and links to the ancestral association we have
with the natural world, where temperatures and light intensity are constantly changing (Image:
Emanuele Naboni)

to English, which uses design now in terms of extending the senses. We can observe
how the quantitative measure of numbers (parametric) and the interpretative, irreg-
ular, and broad qualitative nature of the word “design” are intertwined with each
other. In a way parametric design relates both to the exactness of environmental sets
of ecological, carbon, and people data, as well as on the recognition of local flows,
signs, and pattern reading (Fig. A.2).
You stated, in the beginning, that you were criticizing this kind of directional
research which comes from a rigid manner of thinking where we go from trying to
optimize a certain kind of system by isolating or focusing on it, which contradicts
with the fundamental tenet of ecology where everything is connected to everything.
I liked when you said, “the goal would be to be a doctor caring for the territory.”
That’s a beautiful image that needs further elaboration, but are we designing systems
and redefining the porosity of their boundary conditions to overcome this, let’s say,
scaled thinking or this kind of approach to a particular problem, which emerged
out of site-specific conditions? But then, if we are not able to view a certain local,
let’s say, interaction on a larger scale, it might just simply not work. I understand
your criticism of the current trend to ‘greenify’ certain problems and to want to
76 Epilogue

make cities look like forests. We are not getting any further because we are just
replicating existing systems and bringing them into a new context, but without letting
them actually connect or ‘take root’ there to bring forth new kinds of environments.
So, when we look into computation from a creative point of view, there is a lot of
untouched territory, not only in terms of tool-specific solutions, but stepping a level
above to the macro and addressing what we actually need to solve beyond this kind
of conventional understanding. And that is what Toni and I also touched upon in the
work on this book, to take the discussion of architecture and nature to the next level.
In other words, not merely replicating certain functions and tweaking established
strategies, but rethink our position from a new perspective. I would be curious as to
your speculations about the next steps: How can we combine the competencies and
the pieces of knowledge we have to make that next concrete step to re-envision the
way we want to interact with this planet? Is this the moment where we can use all
this specific knowledge to go beyond optimizing and stimulating sustainable ideals,
to allow ourselves to fundamentally rethink the systems and the questions we have
to ask? I would like to discuss with you a bit further on where you see potential, and
how we can get there. We talked about combining disciplines and all kinds of new
elements and approaches, but still, I have the feeling not much has changed within
the last couple of years. In terms of computation, we just seem to develop more and
more tools.
Right. Thanks so much for summing up my comments and, yes, you are touching
on quite an interesting question. When I consult designers, there is a mind barrier
where design is seen as a multi-domain scientific setting that tackles climate change
and artistic operation at the same time. In my experience, architects are generally posi-
tive about using simulation to solve complex problems related to a building’s struc-
tural design or a bioclimatic issue. Some architects completely rely on computation
as a language informing a style. As a matter of fact, simulation is very rarely adopted
in its full potential as a tool that simultaneously embodies ecologic, climatic, decar-
bonization, circular design, and health means. Simulation serving broad systemic
thinking is not being explored in current architectural practice. The reasons that
hinder this possibility are twofold: a lack of understanding of climate change and
the related breakdowns. Selecting which variables to use and how their relationships
influence ecosystems and human well-being, or other design parameters is a compli-
cated initial step since it sets the edges of the possibility space. This requires consid-
erable knowledge in scientific domains, as well as an understanding of algorithms
if optimization techniques are embodied. The stylistic impact of simulation must be
considered when intended as an integral part of the design inquiry. It became frus-
trating to most architects when they realized that stylistic matters and design freedom
could also be framed by climate change.
I would like to follow up on this. From what you are saying, the traditional archi-
tectural viewpoint still has challenges to understand, for example, process-oriented
systems as design inputs. As outlined through that example, which is probably very
widespread, there is still a very traditional object-oriented approach when it comes
to architecture, versus the kind of process-oriented environment in which we are
living. To recognize this means to interrupt this way of thinking, which connects back
Epilogue 77

to what you also said when you touched upon the flows. These kinds of changing
environments are adapting but still following the same kind of logic, whereas we are
designing fixed entities, which are stable and eventually have a smart environmentally
friendly façade. But that’s all. We are probably looking into a shift of understanding
of what architecture is, when facing the complex problems that confront us. So, we
need to talk about designing processes and not entities.
That is a fantastic definition, yes.
I think it also shows that we probably still have a very human-centric perspective,
and even though we talk of the built environment, we are not used to designing for
insects, designing for trees, designing for nature. So, in some way, we are not used
to building environments. Very likely this is the biggest shift we need to make from
building buildings and cities to building environments and understanding them also
as environments. That requires a much, much broader understanding. That relates to
what we were saying before. We are talking about the porosity of two systems. The set
of rules: one set is related to nature, the other is related to a city. If you look into the
data and the figures, what we are also trying in this joint research is to impose that
as a design driver. So, we are not yet there. I think that is also why these discussions
are so important for us, to see, in a way, the limitations and the potentials, like where
exactly we can tackle this driver for change.
Thanks for putting it in those words; they crystallize the concept a little bit more.
I like this concept of the porosity of domains.
Yes. I think, maybe just to add to this concept of porosity as Pia calls it: If we take
a systemic view really seriously, then this notion of porosity is really key because, in
the end, it’s all about controlling boundary conditions, which define the condition
of exchange. It is all about these questions of how things move from one domain to
the other and one scale to the other. And it is exactly these kinds of different levels
of porosity, I think, that are crucial for this more systemic perspective, which means
to get away from an object-focused perspective into a perspective of exchange and,
yes, flows of porosities in some way.
Sure. And with climate change, porosity will have to be in place: a system able
to modulate inputs and outputs and architecturally manifest such flows. Architecture
interplays with both local and global conditions; it is impacted by climate change
(input) and will impact the local and global climate (output). Closed design systems
curate only inputs or are only able to limit outputs, such as those exacerbating the
climate. A porous system, on the other hand, is surely a possible solution to climate
change.

A.2 Interview with Christophe Girot

In your work as a landscape architect as well as in your academic research, you


sometimes refer to the notion and understanding of the place in relation to the
process-oriented potentials and challenges that play a fundament role as a ‘return to
terrain’. With the rise of performance- and optimization-oriented design, the trend
78 Epilogue

for creating universal solutions has increased. In your research on topology you
point out the importance of the interrelated cohesiveness of things and people. How
do these outlined relationships support a new form of process-oriented design across
scales?
I think the word topology has been reduced to a purely mathematical formula in
the last century, whereas it used to be a much broader philosophical term. It used
to be about thinking about the land and thinking about the surfaces in the land. The
reason why I reintroduced this word topology is because architects use a similar
word, tectonics, to talk about hidden forces within a building, which give rise to
its unique form. These forces are like the force of gravity, or the forces that hold
something together. I was feeling that in landscape architecture, we were missing a
strong abstract term that would bring the understanding of a place together. The idea
is to bring back a consciousness about a place, about how a place is constituted in
its entirety. We have developed a site-specific data gathering methodology, based on
point cloud technology, which starts with the scanning of a site. The pixel points can
be compared to dew drops on the ground. The whole landscape is covered with these
pixels, which define a skin. They do not go deep into the ground, but just define a
surface. This fact is what really convinced me that that topology was the appropriate
term. The goal of topology is not just to be topologists, which we are not. We are
not developing mathematical schemes on continual surfaces. Rather we are actually
trying to bring back wholeness or togetherness in a given place, in a given site. This
is where I think the big difference lies with respect to the question that you were
asking about the modern obsession for optimization and this whole approach to a
world standard in architecture. I think we are completely at the other end of the
spectrum. I believe that each place, each location on earth, has its unique qualities
that give rise to very different phenomena and surface landscapes, or topology. You
could basically argue from a geodesic point of view that from the North Pole to
the Equator, every square kilometer moving down is fundamentally different. And
therefore, the way you approach things, whether it is in the physics of the place, in the
geology of the place, in the climate of the place, in the vegetation, in the population,
in the hydrology, basically means that we use topology as a hypercritical tool to
look at very local conditions, and understand them fully. The problem I had in my
education as an environmental planner and architect, and landscape architect, is that
everything is reduced to a sheet of paper, and we imprint what we see—the entire
world—on a piece of paper. The danger is that not only are we doing this on paper,
but the paper then goes back and imprints itself on the world without correlating to
the reality of the terrain. The computational design revolution that we are looking at
right now is to identify the factors that make the site specific: How do you bring them
together? How do you combine them? And it goes on from terrain conditions, soil
stability, rain, climate, and temperature, all the way to how you put a construction
into the ground. How do you deal with the topology of a place and have it work
together with the proposed design?
Continuing on your findings of the “flat and layered” design approach—relating
as well to McHarg, one might argue that there is a need to rethink the vertical
connection between layers of data. This relates as well to place-based processes
Epilogue 79

and knowledge transfer between domains and questions like, how do we go about
unveiling the hidden knowledge and rules of these processes so that we can design
with them? Within our research and teaching we are speculating on possibilities to
decode the drivers of site-specific systems. How do you see the potentials to bridge
from these, let us say, site-inherent qualities or challenges, towards a new form of
designing which can hold the challenges of the time? And furthermore, how can
your outlined design thinking expand from landscape-oriented questions towards
rethinking what cities of the future can be like? How do these outlined relationships
support a new form of process-oriented design across scales?
The mapping overlay of information, today’s GIS systems, treats data in a system-
atic manner: the data is analyzed, rationalized in schematic form, and in coded and
colored surfaces, to be then overlaid. And then, through what you would call deduc-
tive reasoning, you are able to go through this “layer-sandwich”, to deduct inter-
ventions or areas that are in danger, etc. In comparison to the deductive reasoning
embedded in GIS systems, I have always been more interested in the inductive mode
rather than the deductive mode, and the reason why I am saying I think we can deduce
everything we want from global warming. We have access to all different kinds of
reports telling us, e.g., the change in temperature and possible reasons. What do you
do with these reports at the end of the day? How do you react? I think that we are
at a period now where we really need to act decisively and knowingly. So, coming
to your question, I do not need to tell you that the point cloud revolution is not just
about the settings, it is the suspended pixel point in which you can stack tons more
forms of information than in the McHarg model. I think that the future really lies
precisely in that, how you bring in metadata into the model. How do you actually
filter that metadata, or stack it or not stack it? At ETH Zurich, the team of Prof. Dr.
Grêt-Regamey (https://plus.ethz.ch/) conducts research in the area of hyperspectral
photography, which is photography that goes beyond the visible spectrum and gives
a huge amount of data on the physics of natural phenomena. This relates to sensors
and to give you an example, we are going to put geolocated sensors in the forest, to
link the gathered data with the point cloud model, generating a database. This will
open up questions like: What is the 24 h-cycle in a forest? What happens day and
night, and how do you map these processes? This allows us to get into very precise,
super-informative data about the body of nature. How the body of nature is actually
acting, reacting to different climate factors, different temperatures, different hydro-
logical conditions—a library of geolocated metadata. Currently, many researchers,
working on the Amazon, Borneo, or the Indonesian jungle, are still working mainly
in 2D. They do not even know the cubature of their forests. They will give you just the
square meters, which is absurd, because in the cube is where life happens. It is in the
volume. If you start scanning bits of forests and start looking at the processes that are
actually occurring there in real time, because you have day and night sensors, I think
that is going to create a quantum leap in knowledge and also in possible interven-
tions. I think that is probably one of the next frontiers, and it is truly a computational
challenge because, how do you combine this information? How do you assess the
priorities? Are there priorities? Are things running in parallel that interact or not?
It is basically “landscape medicine”, in a way. It is a bit like the informed imagery
80 Epilogue

for a surgeon before he operates. We can almost imagine that in the near future,
within 10 years from now, with the sensors, with hyperspectral photography, with
the scanning of different parts of our natural environment, we could get to not only
a highly informative level, but also to a level of information that allows to make the
right decisions, the right steps to move forward in this very endangered world. I am
very optimistic. I was always fascinated by the beauty of the thing, but actually, it is
what is not visible, or the hyperspectral, that is the key, partly at least in the natural
world, to really some of the major issues, such as how to maintain nature, how to
reinstate diversity, etc. This is why I believe in the pixel. I think the power of this
topological approach is that each pixel can harbor an incredible set of information,
and then we need to be able to recognize how they interact with one another. This is
the next frontier, and you know that as well as I do, the main obstacle to unleashing
this power is feature recognition. That is to say, the point cloud does not make any
difference between a rock, a tree, a car, and a beer can. It is just a point that appears
in the picture. With feature recognition, we have to train a computer to recognize
entities or patterns, I think this is going to be the next challenge. We are still only in
the Stone Age of this development.
It is very interesting to see the interaction between the typological approach and
the integration of technological advancements. Over a decade ago, the challenge
was the capturing of the environment, and now point cloud models start to be almost
like a knowledge library in which we geolocate different knowledge spaces, in order
to understand how the processes, interact with the site-specific systems. Currently we
are capable, like in the case of digital twins, to go into simulations, aiming for a better
understanding of consequences, beyond the pure simulation of one system. Having
a multisystem simulation tool at our hand, which allows us to run data-informed
predictions, but as well as simulations over time.
Yes, exactly, and I would tend to say the sensor dimension, as it is in real time,
is going to be the key. That is to say, it is not fixed like a McHarg plan, colored
by number, where the information might be outdated and unclear how it has been
recorded. Whereas we talk now about information with specific sensors, like if we
look at the gaseous exchange in a forest between CO2 and oxygen. I am just giving you
an example on forest behavior. Some people are talking about lack of photosynthesis
as being a problem in the near future, too, because of heat gain, etc. The process I
did describe is giving you a real-time assessment of something that is alive. In a way,
it is not completely alive, but it is getting closer to, not just simulating, but reporting
on how the live thing is behaving—it is about forest behavior.
I agree on the overall importance of understanding complex behavior. I was quite
fascinated to hear how you described the use of point clouds in your work. But when
I was listening to you, I thought what you were describing is a lot about the field of
data gathering, that this has clearly evolved compared to the time of McHarg. But I
think we are all not scientists. We are not geologists. We are architects and landscape
architects. So, in some way, I think that we would always like to go beyond collecting
data and transform them into something that is working with these processes, but
goes beyond them. In our explorations, we somehow want to reach an artistic and
cultural level or dimension. So, we are treating data as a scientist would do. The way
Epilogue 81

you have talked about computation sounded more like a rational approach based on
calculation. But I think what Pia and I are trying to do is to think of computation as a
method to deal with this data differently, to connect it and give it a new interpretation,
to allow architecture and landscape architecture to find new meaning. And to me,
this understanding of computation seems to be very close to your understanding of
something like landscape imagination, the idea of bringing another dimension into
the work. I do not know if I am right here; you might disagree.
I was answering Pia very analytically beforehand, because she mentioned
McHarg, and I just wanted to give the example in a world which has 3D, has gone
in real time in terms of analytical power, and that requires computational skills. But
your question, Toni, is about design. And that was always my concern as well: How
do we bring back this approach and this technology to design? And how do we
intervene computationally as well? You probably know the book “Robotic Land-
scapes—Designing the Unfinished” edited by Ilmar Hurkxkens, Fujan Fahmi, and
Ammar (2021). The book was a real challenge for me, because I did not believe in
robotic landscapes at all. I’m not sure I still believe in their existence yet, but it was
a very interesting experiment (note: “Robotic Landscapes—Designing the Unfin-
ished” investigates the use of robot-based construction equipment for large-scale
soil grading in landscape architecture. As landscapes are continuously changing due
to ever-changing environmental conditions, the application of autonomous systems
that respond to the environment rather than perform predefined and static earthwork
is of particular interest in this field).
The experiment was very interesting, because I realized that we were in a triangle.
On one side, you have environmental forces. These forces, whether you like it or not,
are definitely in the landscape. The wind is there, the water is there, and the forces
are there. The unpredictable dimension of nature is there. And, on the other side, we
have the whole palette of computational tools and robotic guiding systems. Fabio
Gramazio (https://gramaziokohler.arch.ethz.ch/) is really into developing technolog-
ical solutions, not teaching design or styles. He is saying there is a machine language
that can start generating options that have an independent logic. And so, we were
taking these two forces, we can say computational design forces and environmental
forces, and trying to bring them together. And the architect or the student that was
stuck in the middle of these two things studied art history. They were not equipped
to deal with this confrontation, with this clashing of realities. And the result of the
dilemma of how do you finally decide or design with the information was actually
very manual and intuitive. What is really fascinating, Ian McHarg initially devel-
oped these tools many years ago. We have been developing tools of design directly
in the point cloud model. So, we do not use the point cloud model merely as an
analytical library or a cemetery of information, but we actually use it as a terrain
for modification. There is this one example, which I think is really quite telling. We
have not shown it that much outside the school yet. It was on the Gürbe River, in the
Bern Oberland, where for the last 150 years they have had these huge, phenomenal,
and very violent debris flows. Entire pieces of mountainside would come down in
the river and behave like water. 15-ton stones flow like water down into the valley.
They crash through villages and destroy everything. And so, over the last 150 years,
82 Epilogue

engineers have built 120 check dams. They built dams all the way up the valley,
to stop the power, the violence of the phenomenon. We said, we will get rid of all
the dams, and we will just modulate the terrain to receive the event. The students
managed within one semester with the computational tools at their disposal to design
the project. The integrated tool Docofossor (a terrain modeling plugin for Rhino 6 +
Grasshopper) developed by Mathias Bernhard and Ilmar Hurkxkens (https://www.
food4rhino.com/en/app/docofossor) was a totally un-architectural tool. Nothing is
flat. It was always about inclines and slopes. And the students started tweaking piles
of river material, and shaping the river material in such a form that it could take
the event and hold it, hold it within the given frameworks. What I am trying to say
is, we are only at the beginning of this process of gestaltung or design. What was
interesting about this, and I am referring to the triangle I spoke of previously, the
architects and the students were able to decide how to shape things and how to posi-
tion them. The computer could not do that; it could only increase the number of
possible variations. Within the feedback loop we would test the design, we would
see the water behavior, and we used programs from the WSL—Swiss Federal Insti-
tute for Forest, Snow and Landscape Research WSL (https://www.wsl.ch/en/index.
html) to simulate debris flow and little by little we observed and reacted by tweaking
the design. Between the violent natural event and the robotic shaping, this tweaking
was a human act. It was what I would call a design act, which, finally after 20 runs,
mastered the problem entirely. This was very satisfactory, but you could also say it
was not highly aesthetic, that it was not designed yet, it just dealt with the problem.
But if you look to the future, it is going to be about coastal protection and river
protection. We are in the middle of it. I was not totally convinced by robotic design
methodology. The robots in the scheme were like dogs in a little house. They wait
for the event, and then they go out and they do their thing, and then they go back in
their house and wait for the next event. So, the robot in the little house replaces the
concrete dam. The robot is the watchdog, in a way, just coming out and dealing, in
real time, to control the situation when the event is happening. It is an experiment.
I am not saying that our design studio approach is the answer to everything, but it
was non-analytical. This was pure inductive design. It goes back to what I was saying.
The design was not deductive, it was inductive. You try a shape. It is a catastrophe.
You do it again. You incline. And what the students realized is, by just working very
slightly on the inclination of a slope, you can completely change the physics of the
river. The end result was really pretty amazing—like “haute couture” (Fig. A.3).
I think this is a very good example, because it very much illustrates how we
understand computation as an inductive process of playing, in your case, the forces
of the landscape to produce, yes, a ‘gestaltung’, if you want to call it that. In your
case, of course, the goal was to redirect the flow, but in some way it is still a design
that has a purpose, but is working with natural forces. And I think working with the
tool is maybe about intuition and trial and error. I think it describes very well how we
understand computation not necessarily as an analytic rational form of working, but
rather as an explorative tool that works with data in this inductive form of designing.
In that sense, I think we have pretty much the same understanding of how to work
with these things.
Epilogue 83

Fig. A.3 Iterative process diagram describing the data-informed robotic design process utilizing
the sand-box. The process consists of informed topological performance studies (upper row), which
are simultaneously used for the automated generation of corresponding 3D models (in the middle
row) and the graphical representation of the script that informs the robotic arm (lower row). Student
project by Ladina Ramming and Thorben Westerhuys (Image: Christophe Girot/ETH Zurich)

We gave students computational design tools, which they learned in a month, and
then were able to tackle this complex design task. This is phenomenal. They worked
on major and very real environmental problems. Of course, in the framework of the
studio, it would probably cost only 10% of what it would costs with concrete. The
engineers will be furious with this because they basically will lose their jobs. As soon
as we condition the robots to do their work, then there are no more engineers there,
so there are a lot of lobbies resisting this change. But yes, we have started developing
these very powerful design tools.
What drives the research of Toni and me is the interest in the activation of cities
as a generator for balanced habitats. We are not questioning the static compo-
nent of architecture, but looking into ways, similar to your examples of the Robotic
Landscape projects, into processes informed by nature and flows, which through the
combination of topological changes as well as site-specific forces, create a design
which allows for adaption over time. We are looking into computationally informed
methods, how we can work with the site-specific systems, which might be environ-
mental, as they might be societal or even political. How these processes can be
decoded and hopefully also allow us to enter a discussion about what is actually
architecture or a city of the future. And I think this is where the elements of our
conversation nicely converge. Looking back to the beginning of our interview, I have
criticized the performance- and optimization-based approach which is very different
because it is less oriented to the general question about design. And this is where
we are trying to fuse these digital, physical, and also biological worlds through a
new way of computational design thinking. You have been working on a huge range
of projects focusing on process over time. Many of these cutting-edge projects deal
with blurring the edges between architecture and landscape architecture. Is there a
84 Epilogue

trajectory of convergence that you have observed, or do you still see these separations
of fields?
There is a general tendency to enclose. I will give you the example, the “Minergie
fashion” (international eco-label for buildings, based on low energy consumption)
in architecture, putting three pullovers on a house, so that someone can walk around
in their bikini in January in front of their window. This idea of enclosing the human
shell completely into a self-sustained environment that is almost autistic to the outside
world. I think the key to what you are talking about, if we are really going to talk
about the urban scale, is what we call urban physics. And it is actually one of the
missing links in the chain. You have these guys reporting on the climate right now,
the G8, but they are working on a stratospheric scale. They are working on the scale
of a continent. They are working on the scale of a whole country. They are working
on the scale of a whole region. But if you ask them about a city and the specifics of
microclimatic in a city: zero, nothing. And then you have all these people saying,
yes, we are going to plant more trees, and it is going to be cool, and everything is fine.
In the meantime, the architects are building these boxes that are completely closed.
You cannot even open the window anymore. The real problem today is precisely, as
you are saying, the separation. That is why I think the word topology comes again
correctly into use. I have been working a lot with Prof. Dr. Jan Carmeliet (https://car
meliet.ethz.ch/), one of the pioneers in urban physics. I would tend to believe that,
before we look at architecture, we should look at the scale of a city. We should look
at the topography of a city. We should look at the heat spots and the cold spots in
a city. We should look as well at the underground and its infrastructure. What I am
saying is that urban physics is really important. And again, I could use exactly the
same language I was using for the forest. You need to map the urban physics of a
town on 24-h cycles, probably on four seasons, and the day/night, the way things
move in one direction. The physics of a city aren’t really being studied. I would like
to point towards a missing link, and the missing link is, again, looking at the city as
a body. And if it is breathing, there is a lot that we can do.
The first thing I would say is, architecture needs to get rid of this Minergie styro-
foam overcoat and open up. There needs to be that exchange, because otherwise, if
we start creating parallel worlds with boxes doing one thing and trees doing another,
where is the coherence? Where is the topological coherence between all these deci-
sions? I believe very strongly, and it is a domain in its infancy, in the importance
of urban physics. It is a computational paradise for whoever likes computational
complexity. It is really looking at how the city is an organism, very much like a
forest, breathes, lives, day, night, etc. Very soon, we are going to have to act strongly
on cities. Maybe every fourth street will be carless. It will be very different. But
then, how do you organize that? How do you understand the mechanics, specifi-
cally the physical mechanics of what you are doing? It varies from place to place.
We thought, naively, of doing a studio in Singapore, where we were going to cool
the city center. The urban physics, the temperate climates under the Equator were
completely different than here. Again, it goes back to what I was saying right at
the beginning. Every square kilometer, from the North Pole to the Equator, has a
different set of rules. Cities look alike, but they have very different parameters. And
Epilogue 85

Fig. A.4 The existing cargo port in Tanjong Pagar will make way for a new expansion of the central
business district of Singapore. The Rail Corridor with its terminus station could be integrated into
the strategic use of the open spaces to dissipate the urban heat island effect. At street level, the airflow
improves thermal comfort. Designs by ETH and SUTD students applying the design method by
Philipp Urech, simulations in OpenFOAM by M.O. Mughal (Image: Philipp Urech/ETH Zurich)

I am very optimistic. I think we should really develop that. I do not know which
school will do that, but it will be absolutely outstanding if you did (Fig. A.4).
I think the example that you brought up, with the critical view on green energy
and all these attitudes that we discussed, demonstrate how sustainability is broadly
viewed, which is quite an important aspect. Because the way we discuss sustainability
nowadays is very much with an attitude of scarcity. It is always reducing and talking
about the reduction of impact on some form of environment, with the result, like
you said, of wrapping the building more and more, to be totally isolated. In some
way, it is also a sign of a form of thinking, where the manmade world and what you
call the natural world are still seen as two different spheres, not working together.
The example that you brought up with urban space and the need to understand its
physics is in some way similar to the idea of understanding the physics in landscape.
This is pointing to the question that we will probably need to resolve this opposition
between a manmade and a natural world. And in some way understand more the
manmade urban space, the architecture as part of a natural system in some way.
Would you agree with this, that this is in some way still a major challenge, to be able
to fully understand, accept, and conceptualize differently than we have in the past as
architects and landscape architects?
I think everybody would profit from such a vision. I agree with you. I think we do
not have a choice. We are going to have to open up to the world around us, to make
us much more aware of everything that is going on, and help us look at the extremes.
But it is really a question of comfort, too: a definition of consumerism comfort. With
86 Epilogue

the energy crisis we are heading for right now, which is going to be a major energy
crisis and a major turn, some people may end up just having, like in previous times,
one room that is really well heated and then others that are a bit cooler. And we
are going back to a more simple and pragmatic way of living, more in contact with
reality. So, it goes back to that whole question of, is BIM the last word on the city,
on architecture? These solutions need to go way beyond BIM. We need to think in
bigger entities, in bigger ensembles. And, of course, as Toni was saying, we need to
bring the inside and the outside much more together. We do not have a choice.
I guess where I do see a strong power in having a data-informed way of simulating
certain kinds of climate actions, is that we can actually show the impact from local,
small changes within the global setting. And that is what is needed in order to inform
real changes in re-thinking the cities. This relates as well to the need of shifting our
attitude. It is about going from a very personal or individualistic view towards being
a part of a bigger, we can call it system, or you called it, “the city as an organism.”
That is probably the only way we can respond in a meaningful manner to the complex
challenges we are facing, and also as you said, to take advantage of rethinking site-
specific interaction with natural processes. I know that you were critical about the
robotic experiment in the beginning, but I think this will just be the way in which it
will go. And it will probably lead into similar applications within cities.
I learned a lot. It was technical, but I learned a lot from this excursion into robotics.
In the end, it was a great experiment.
I think when we started at the beginning discussing about topology, the way you
were talking about this reminded me very much of the original notion of topology,
which was called analysis situs. I think that that fits much better to what you are
talking about. Leibniz was the first using this notion in his investigations into the
foundation of geometry. For him the concept of situation, Latin situs, is about the
position of an object in space relative to other objects. Thus, from the very beginning
of Leibniz’s investigations, it is considered a relational predicate and source of
Leibniz’ theory of a relational space. In this sense, analysis situs is a kind of almost
philosophical exploration of a site. Something that will go much further than a purely
technical approach. I had the feeling your understanding of topology relates much
more with this original notion.
I agree, I like your word in situ. Yes, topology in situ. If you look at Leonhard
Euler’s Seven Bridges of Königsberg, he took the seven bridges in Königsberg as
a physical reality, to demonstrate an abstract process. Which is, you cannot come
back on the island if you only cross the bridge once. It seems the reality is driven by
formulae or these other topologies or mathematical topologies. And I think this idea
of bringing topology back in situ would be closing the circle with Euler, and bringing
back the abstract formula to a more concrete reality. And just to recap that, so that
you do not understand me wrongly, for urban physics, for instance, the problem is
the information on climate is not at the right scale. The global information is not
the local information. As in any scientific venture or research, the scale is the scale.
If you are at the microscopic scale, you are not at the nanoscale. If you are at the
one-to-one scale of architecture, you are not at the scale of the territory. What I am
saying is, to each scale its method. And the only thing I wanted to say is, the city will
Epilogue 87

have sensors too, and will end up analytically understanding a story, which will be
very different than the common discussion we are having right now. The problem is
that most decision processes and urban thinking are still lying flat on a map. People
do not even understand the terrain, do not even understand how air moves down or
up a slope, how water moves, etc. All I can say is, the analogy I used about the forest
is exactly the same. To me, it is the same thing. We should look at the forest and the
city as the same kind of entity: as an organism. It needs to be analyzed, censored,
but also inductively modified.
These are beautiful words to close this conversation.

You might also like