Structuralism and Form in Literature and Biology: Critiquing Genetic Manipulation 3031477383, 9783031477386

The book considers biology in parallel with philosophical structuralism in order to argue that notions of form in the or

123 109 8MB

English Pages 239 [253] Year 2024

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface
Acknowledgements
Contents
List of Figures
Chapter 1: Structuralism, Vitalism, and Bioengineering
Genetic Manipulation Technologies Enter Under the Radar
Dreams of Genetic Manipulation
Gene Technologies Have Potential Benefits
Focus Is Placed on Practical Risks
The Code of Life and Bioengineering
Structural Relations and Vitalism
Structuralism Across Disciplines
Formalism Returns with the Modernist Movement
Indivisible Form
Form Is Conflated with Vitalism
Cassirer’s Critique of Both Vitalism and Reductionism
Boundary of Life/Non-life
References
Chapter 2: The Debate: Cuvier and Geoffroy
Controversies in Early French Biology
Background to the Debate
Rational Order of Eighteenth-Century Natural History
Transition to Modern Biology
Les embranchements
Geoffroy Put Morphological Structure First
Form as Homology
The Debate at the Academie des Sciences
Le transformisme
References
Chapter 3: Setting the Stage for Evolutionary Theory
Paradoxes Arise in Pre-Darwinian Theory
Demise of Formalist Biology
An Atomistic Paradigm
Gene and Function
Systems Biology Entrenches the Reductionist Approach
Functionality Attributed to Material Components
The Road to a Historicist Biology
Bringing Back the Formalist Ideas of Early French Zoologists
References
Chapter 4: Adaptationism and the Author
Adaptationism: A Functionalist Paradigm
The Adaptationist Account
Darwin’s Synthesis, Adaptation, and the Gene
Early Formalist Ideas in Biology
The Author, Romanticism, and German Archetypes
New Criticism
The ‘Author’ in Biology
Humanist Causes
Questioning Current Accounts of Evolution
The Organism as a Unit
Structural Transformation, Not Malleability
References
Chapter 5: The Relational Turn
Autonomy in Literature and Biology
Biology as a Science in Its Own Right
Formalists Turn the Tables on Functionalism: Literary Formalism in Russia
Relational Biology: Life as a System
An Immanent Organisation
Critiques of Modern Systems Biology
Infinite Regress
Non-computability
An Alternative View
References
Chapter 6: Prague Structuralism and the Poetic Function
Structuralists Question Neo-Darwinism and Genetic Manipulation
Langue: Saussure’s System
Impact of Structuralism
Limitations of Structuralism
The Prague Linguistics Circle
The Metaphor-Metonym Poles
The Poetic Function
Literary Systems as Historical
Revising the Synchrony/Diachrony Distinction
Connection to Extra-Literary Factors
Jakobson’s Studies on Aphasia
The Flexible Phenotype: The Role of Context in Biology
Adaptation, Adaptivity and Phenotypic Plasticity
References
Chapter 7: Immanent Evolution
Literary Evolution as an Internal Process
Modification and Evolution
Transformism and the Challenge to Fixism
Evolution and Orthogenesis
Complexity
Transformation of Structure: Tynianov and Geoffroy
Langue that Incorporates History
The Organism as a Transformative Structure
Morphogenesis
Self-Organisation and Homology
Importance of Stable Traits
Gene Are Ancient Structures
References
Chapter 8: Phenotype then Gene
Evolution by Combining Traits
Going Further Than Constraints and Self-Assembly
Sudden Evolutionary Shifts Proposed by Biologists: Mutationism
Punctuated Equilibria
Modular Systems
Radical Phenotypic Transformations
von Baer’s Rule
Phenotypic Accommodation
Genetic Accommodation
Internal Transformation and Deformation
A Revival of Saltation
Is Homoplasy Really an Instance of Recurrence?
References
Chapter 9: The Interpreting Organism
The Affective Fallacy
Biosemiotics
The ‘Umwelten’ and Biosemiotics
Differentiating Coding and Peircean Semiotics
Divergence from Structuralism
A Revised Structuralism
Ideology and Realism
Realist Literature
A Decentred Biology
References
Chapter 10: An Ecological Context
Questions Raised about Approaching Living Systems as Material for a Humanist ‘Cause’
Modification: An External Cause
Discourse of Modern Agriculture
Context of Environment
Quantitative Breeding: Contributions from Population Biology
References
References
Index
Recommend Papers

Structuralism and Form in Literature and Biology: Critiquing Genetic Manipulation
 3031477383, 9783031477386

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Structuralism and Form in Literature and Biology Critiquing Genetic Manipulation Peter McMahon

Structuralism and Form in Literature and Biology

Peter McMahon

Structuralism and Form in Literature and Biology Critiquing Genetic Manipulation

Peter McMahon Eltham North, VIC, Australia

ISBN 978-3-031-47738-6    ISBN 978-3-031-47739-3 (eBook) https://doi.org/10.1007/978-3-031-47739-3 © The Editor(s) (if applicable) and The Author(s), under exclusive licence to Springer Nature Switzerland AG 2024 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 Palgrave Macmillan imprint is published by the registered company Springer Nature Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland Paper in this product is recyclable.

To Wilhelmina

Preface

To navigate by a landmark tied to your own ship’s head is ultimately impossible. Donald M. McKay1

Genetic manipulation (also known as engineering or modification) technologies are implicit to modern biology. As well as potential benefits, they raise concerns regarding unintended consequences of their applications. Numerous commentators have addressed these concerns. But what is significant, as opposed to the consequences of genetic manipulation applications (intended or not), is the seeming dearth of questions about the unproblematic acceptance of these technologies. The introductory chapter of this book quotes Wes Jackson saying, as far back as the early 1980s, how genetic manipulation technologies have entered mainstream science under the radar. These technologies are now an integral and unquestioned part of modern biology and scientific endeavours in the medical, agricultural, and medical sectors, including more recent aims to build organisms through artificial life simulation and synthetic biology. Some of these technologies will be useful, while potential problems have also been uncovered by activists, concerned citizens, and scientists. Yet the underlying discourse that appears to give these technologies the carte blanche to forge ahead is largely unchallenged. Some have, however, questioned the premises of genetic engineering (or modification); they include biologists who have reservations about the currently accepted neo-Darwinian paradigm, including the modern 1

 Ciba Foundation, Man and His Future 1963, London, p. 286, quoted by Hans Jonas.

vii

viii 

PREFACE

synthesis which brought Darwinian theory and genetics together. Their ideas and views have contributed to the themes developed in this book. As an undergraduate at Sussex University, I took a course unit on developmental biology taught by two of these biologists, Gerry Webster and Brian Goodwin. I would like to say that their lectures inspired me to continue reading about structuralist interpretations of biological phenomena, but I seem to remember that at the time I was much more interested in other parts of the course—microbial genetics with Oliver Darlington and Sheila Maynard-Smith, evolution of animal behaviour with Paul Harvey, plant-­ water relations with Prof. James Sutcliffe, among others—so that much of the developmental biology module went right over my head. However, reading their papers much later was inspiring and a spur to learn more about the history of ideas on evolution, including structuralist approaches. A further resource has been my grandfather’s book The Divine Proportion, long on the family bookshelves, which explored patterns of mathematical form in nature, both organic and inorganic. An ordained Methodist, Ted Huntley took a ‘design’ approach to natural phenomena, but nevertheless raised perennial themes that are relevant to our understanding of living forms (or more accurately our lack of understanding). Structuralism, beginning with the redefinition of language as a system by Ferdinand de Saussure in the early twentieth century, became the fashionable ‘agora’ of theory in the humanities, reaching its heyday in the 1960s. Structuralist biology, although marginalised, even denigrated, in comparison with the high structuralism of the humanities, is derived from similar themes to those in the field of linguistics. Ernst Cassirer, Jean Piaget, and others have pointed to common structuralist themes across disciplines, including biology. My aim in this book is to identify some of these common themes, especially in formalist literary theory. By doing this, the work of literary theorists might shed light on alternative approaches to biology. Structuralist biology presents a critique of the (post)modern attack on the organism and its reconstitution as an informational script, available for patenting and commercial control. It comes under the umbrella of theoretical biology, seen as one approach to a better understanding of the organism. This is because, as demonstrated by formalism in biology or literature, reducing the organism or a literary text to its components is not enough to reach an understanding of its workings, its raison d’être. In the postmodern world, ‘theory’ has been systematically denigrated. Yet in order to develop a critique of modern biological practice, especially genetic

 PREFACE 

ix

manipulation, a return to the structuralist and poststructuralist ideas that rode in on the modernist wave is needed. Formalism (and structuralism) cannot be separated from the sea-change that drove the modernist movement in art in the late nineteenth and early twentieth centuries. Some biologists returned to formalist ideas in the modernist period, producing a flurry of embryological studies and attempts to demonstrate Lamarckism as an evolutionary force. Efforts to confirm Lamarck’s ideas on the inheritance of acquired characters and endorsement of his theme of evolution as a drive to increasing complexity did not yield many results, but some of these ideas, along with biological structuralism, have gained more ground recently. With developments in relational biology and evolutionary developmental biology (evo-devo) that endorse a structuralist account of living processes, the organism is increasingly portrayed as an interacting set of relations, a system that determines, in part, the function of its components. Formalist and structuralist themes applied across different disciplines have helped to paint a new picture of the organism. Criticisms of genetic technologies based on the integrity of the organism and species are perhaps expressed most strongly by exponents of structuralist biology. They have tried to develop secular arguments proposing that the organism is more than its collective DNA and related functions. They promote the organism and its taxonomy as the basic unit of biology and, in turn, reject the premise that a central genetic agency informs all life. As opposed to both Lamarckian and adaptationist ideas in biology, structuralists reject scenarios such as ‘need creates structure’, and that the gene directs organism function. This translates into a critique of the wholesale promotion of genetic manipulation as being the most promising route to meeting agricultural productivity targets, managing the pests and diseases that cause heavy losses and dealing with environmental crises. Therefore, structuralism could provide the basis for a priori critiques of the explosive rise of genetic technologies, in addition to their potential consequences. In the discussion presented here, such critiques are directed not to genetic technologies per se but rather, firstly, to the role of genetic technologies in decontextualizing the organism from its ecological relations and, secondly, to the manipulation of germ line (reproductive) cells. Such manipulated changes in DNA sequences are inherited and permanent. Ethical issues are raised in relation to humans as to whether some gene therapies might be applied to the germ line and whether this could be a ‘slippery slope’ leading to the development of designer babies.

x 

PREFACE

However, regarding the application of irretrievable genetic technologies, the rest of the biosphere (the non-human world) is considered a zone of no limits. The purpose of this discussion is to question applications of genetic manipulation that are specifically aimed at germ line (permanent and heritable) changes (outside of the lab), not molecular technologies per se, which are useful as research tools and in other applications such as somatic gene therapy. In agriculture, the identification of quantitative trait loci (QTLs) and marker-assisted selection (MAS) are important molecular applications, for example. The basic premise behind this discussion is that organisms (that make up the natural world) have some autonomy and should be treated accordingly as beings that could aid human life rather than being reconstituted by it. Eltham North, VIC, Australia

Peter McMahon

Acknowledgements

I would like to express my gratitude to my former supervisors, John Anderson and Philip Keane, for their readiness to share their knowledge and ideas on the possibilities of an integrated (ecological) agriculture. Thanks to Palgrave Macmillan editor Robin James for her guidance. On the ethics of genetic technologies, I learned a great deal from Bob Phelps while working at the Australian Conservation Foundation. Thanks also to Sheena, Andrew and Adriana McMahon, and Chang Insun, for discussions and shared ideas.

xi

Contents

1 Structuralism,  Vitalism, and Bioengineering  1 Genetic Manipulation Technologies Enter Under the Radar   1 Dreams of Genetic Manipulation   3 Gene Technologies Have Potential Benefits   6 Focus Is Placed on Practical Risks   8 The Code of Life and Bioengineering   9 Structural Relations and Vitalism  11 Structuralism Across Disciplines  12 Formalism Returns with the Modernist Movement  14 Indivisible Form  17 Form Is Conflated with Vitalism  18 Cassirer’s Critique of Both Vitalism and Reductionism  21 Boundary of Life/Non-life  23 References  25 2 The  Debate: Cuvier and Geoffroy 29 Controversies in Early French Biology  29 Background to the Debate  30 Rational Order of Eighteenth-Century Natural History  32 Transition to Modern Biology  34 Les embranchements  36 Geoffroy Put Morphological Structure First  40 Form as Homology  42

xiii

xiv 

Contents

The Debate at the Academie des Sciences  44 Le transformisme  45 References  48 3 Setting  the Stage for Evolutionary Theory 49 Paradoxes Arise in Pre-Darwinian Theory  49 Demise of Formalist Biology  52 An Atomistic Paradigm  53 Gene and Function  56 Systems Biology Entrenches the Reductionist Approach  58 Functionality Attributed to Material Components  60 The Road to a Historicist Biology  61 Bringing Back the Formalist Ideas of Early French Zoologists  64 References  65 4 Adaptationism  and the Author 67 Adaptationism: A Functionalist Paradigm  67 The Adaptationist Account  69 Darwin’s Synthesis, Adaptation, and the Gene  72 Early Formalist Ideas in Biology  75 The Author, Romanticism, and German Archetypes  77 New Criticism  78 The ‘Author’ in Biology  80 Humanist Causes  82 Questioning Current Accounts of Evolution  83 The Organism as a Unit  84 Structural Transformation, Not Malleability  87 References  89 5 The Relational Turn 93 Autonomy in Literature and Biology  93 Biology as a Science in Its Own Right  94 Formalists Turn the Tables on Functionalism: Literary Formalism in Russia  96

 Contents 

xv

Relational Biology: Life as a System 101 An Immanent Organisation 103 Critiques of Modern Systems Biology 105 Infinite Regress 107 Non-computability 109 An Alternative View 110 References 111 6 Prague  Structuralism and the Poetic Function115 Structuralists Question Neo-Darwinism and Genetic Manipulation 115 Langue: Saussure’s System 117 Impact of Structuralism 120 Limitations of Structuralism 122 The Prague Linguistics Circle 123 The Metaphor-Metonym Poles 124 The Poetic Function 126 Literary Systems as Historical 127 Revising the Synchrony/Diachrony Distinction 129 Connection to Extra-Literary Factors 129 Jakobson’s Studies on Aphasia 131 The Flexible Phenotype: The Role of Context in Biology 133 Adaptation, Adaptivity and Phenotypic Plasticity 135 References 137 7 Immanent Evolution139 Literary Evolution as an Internal Process 139 Modification and Evolution 141 Transformism and the Challenge to Fixism 143 Evolution and Orthogenesis 145 Complexity 146 Transformation of Structure: Tynianov and Geoffroy 148 Langue that Incorporates History 150 The Organism as a Transformative Structure 151 Morphogenesis 154

xvi 

Contents

Self-Organisation and Homology 155 Importance of Stable Traits 157 Gene Are Ancient Structures 158 References 160 8 Phenotype then Gene163 Evolution by Combining Traits 163 Going Further Than Constraints and Self-Assembly 166 Sudden Evolutionary Shifts Proposed by Biologists: Mutationism 166 Punctuated Equilibria 168 Modular Systems 169 Radical Phenotypic Transformations 172 von Baer’s Rule 172 Phenotypic Accommodation 174 Genetic Accommodation 175 Internal Transformation and Deformation 179 A Revival of Saltation 181 Is Homoplasy Really an Instance of Recurrence? 183 References 185 9 The Interpreting Organism187 The Affective Fallacy 187 Biosemiotics 189 The ‘Umwelten’ and Biosemiotics 189 Differentiating Coding and Peircean Semiotics 194 Divergence from Structuralism 195 A Revised Structuralism 196 Ideology and Realism 199 Realist Literature 200 A Decentred Biology 202 References 205 10 An Ecological Context207 Questions Raised about Approaching Living Systems as Material for a Humanist ‘Cause’ 207 Modification: An External Cause 209

 Contents 

xvii

Discourse of Modern Agriculture 211 Context of Environment 213 Quantitative Breeding: Contributions from Population Biology 214 References 219 References221 Index233

List of Figures

Fig. 1.1

Fig. 2.1

Fig. 2.2 Fig. 2.3 Fig. 2.4 Fig. 3.1 Fig. 4.1 Fig. 5.1 Fig. 5.2 Fig. 5.3

An example of D’Arcy Thompson’s proposal of homology based on transformation: the sunfish on the right (b) could be considered a deformation of Diodon, the porcupine fish, left (a) as homologous points on each fish species have similar coordinates on the deformed grid. (Creative Commons, Wikimedia)19 The pentadactyl limb: top: salamander, toad, crocodile; bottom: bat, whale, mole, human. After Wilhelm Leche (1909) in Man, Origin, and Evolution. Available in the public domain. https://doi.org/10.5048/BIO-­C.2013.3.f1 35 Georges Cuvier holding a fossil of a fish (Wikipedia Commons) 37 A page from Animal Kingdom by Georges Cuvier 38 Etienne Geoffroy St Hilaire (Wikipedia Commons) 41 Sussex biologists in 1968: (left to right) Jonathan Cooke, Gerry Webster, and Brian Goodwin. (Source: This is Sussex, April 16, 2021) 62 The moth Gastroparcha padale mimics a dead leaf. (Vaishak Kallore, Creative Commons) 70 Yuri Tynianov (1894–1943) (Creative Commons) 97 The mathematical biologist Nicolas Rashevsky (1899–1972). (Source: Alchetron.com) 102 A sketch of Rosen’s M/R system. Metabolites S and T are converted to the product ST, catalysed by STU, one of the two catalysts in the system, while an additional metabolite, U, is used for the replacement (or repair) system to regenerate the catalysts (which have a limited lifespan and are subject to decay). The system is materially and thermodynamically open, xix

xx 

List of Figures

Fig. 6.1

Fig. 6.2 Fig. 7.1

Fig. 7.2

Fig. 7.3

Fig. 8.1

Fig. 8.2

but according to Rosen, closed to efficient causation. (Note: Catalyst 1 has two functions—multi-­functionality is postulated to be necessary for metabolic closure. Although the catalysts are shown separately here for simplicity, a fuller picture would be provided if they are viewed as forming intermediaries, or complexes which partake in reactions—see Cardenas et al. (2010)) 103 Saussure’s paradigmatic/syntagmatic axes interpreted by Jakobson as two poles in language: metaphor and metonym. The axis of selection (paradigm, metaphor) is associated with possibilities (a static structure), and the axis of contiguity or combination (syntagm, metonym) with sentences that unfold in actual time 120 Circumnutation, the term given by Charles Darwin to the growth of plants in the dark as they search for a light source. (Isabel Fiorello et al., Creative Commons) 134 Lamarck’s evolutionary theory proposed a vital drive to higher levels of complexity, plus diversification at each stage as individual organisms adapt themselves to changing conditions (Ian Alexander, Creative Commons attribution) 144 Embryonic development in different species (left to right: fish, salamander, turtle, chicken, pig, dog, cow, rabbit, human) in Ernst Haeckel’s 1877 Anthropogenie, 3rd ed. (APS Museum, Flickr.com). Gill structures are present in all species at an early stage of development 155 Alternative splicing of gene sequences: the exons (coloured sections) of the RNA transcript of the DNA sequence (top) are spliced together, but they can be edited in different ways, expressing different proteins. In the version shown at bottom right, the middle section has been omitted (Creative Commons)159 A development pathway showing early and later stages. Changes to genes linked to earlier branches (unshaded) would have major and deleterious effects; therefore, modern evolutionary theory maintains only changes at the later stages (shaded) would be feasible—shown, in this case, at bottom right. (After John Maynard-Smith)173 Genetic accommodation: the phenotype (P) can be converted from form A into a new phenotype (B) by changing an environmental parameter (E), in this case by increasing temperature. (i) The slope of the line with the arrow indicates phenotypic plasticity; (ii) at the higher temperature conditions, the alternative phenotype (B) is selected replacing phenotype

  List of Figures 

xxi

A; (iii) either the genes linked to trait B become fixed so now the original environmental condition (the lower temperature) no longer results in phenotype A, but maintains B and plasticity has been lost (1), a case of genetic assimilation, or plasticity is not lost and the lower temperature results in a reversion to phenotype A (2). (Adapted from Pigliucci et al. 2006)178 Fig. 8.3 Camera eyes have evolved independently: the vertebrate eye (left) and the octopus eye (right). (1) Retina; (2) Neurones; (3) Optic nerve; (4) Blind spot (in vertebrates). These are not homological structures and the two taxonomic groups are unrelated. (Wikimedia, Creative Commons) 184 Fig. 9.1 Umwelt as a cybernetic system (public domain, University of Hamburg) 190 Fig. 9.2 Thomas Sebeok with his parents in Hungary c.1924 (public domain)194 Fig. 10.1 Traditional selection and breeding manipulate organisms for agriculture, fermentation in wine-making and other activities, working with whole organisms, which act as ‘gatekeepers’ to changes, while modern technologies employ direct manipulation, particularly genetic transformation technologies, bypassing the determining relations of the whole organism 216

CHAPTER 1

Structuralism, Vitalism, and Bioengineering

Genetic Manipulation Technologies Enter Under the Radar Writing for The Ecologist in the 1980s, Wes Jackson commented on how biology had undergone a quiet transformation: “[A]lmost without notice, the era of discovery moved smoothly into the era of manipulation until, almost suddenly, we had new household words and phrases such as ‘gene splicing’, ‘gene stitching’ and ‘DNA surgery’” (Jackson 1984, 120). Genetic manipulation (and bioengineering) is part and parcel of current biology. As both an aspiration and a practice, it cannot be separated from the ethos of (post)modern trends in thought. The promises of genetic manipulation are outcomes of the utopianesque approach taken by modern science. Promised benefits of a science of manipulating life were first formulated with Renaissance utopian thought in the sixteenth century, while similar promises now have driven neo-liberalist policies on commodification and patenting of life forms. These policies have been dominant since the 1980s, especially after the fall of the Berlin Wall, accelerating biology into an era of manipulation (Jackson 1984; Cooper 2008). How could genetic manipulation or engineering be defined? Anne Chapman offers this definition: “I take genetic engineering to be the practice which seeks to produce altered organisms through direct intervention at the level of the DNA of the organism, by means of laboratory methods © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_1

1

2 

P. MCMAHON

such as recombinant DNA technology” (Chapman 2005). The term ‘bioengineering’ is somewhat broader but applied to food products amounts to the same thing. The United States Department of Agriculture standard of disclosure, therefore, “defines bioengineered foods as those that contain detectable genetic material that has been modified through certain lab techniques and cannot be created through conventional breeding or found in nature” (https://www.ams.usda.gov/rules-­regulations/be). In human society, ethical issues are raised in regard to genetic manipulation, since modification of germline cells would be carried on to succeeding generations. For example, consent to genetic ‘improvements’ or changes cannot be given by children yet to be born. These ethical issues have been broached by philosophers such as Jurgen Habermas (Habermas 2003), and remain of greatest concern to the future of humanity. However, the focus of this book is on genetic manipulation in non-human spheres, such as agriculture and environmental management. In these fields, focus is placed on the potential impacts of genetic technologies, such as environmental or nutritional (health) consequences, rather than more basic ethical considerations. In fact, after a flurry of concern in the 1970s, when the genetic manipulation of organisms was first demonstrated and numerous committee meetings were held on the ethics of the new technology, genetic manipulation has become largely ‘ethics-free’. The quiet transformation Jackson referred to means genetic manipulation has entered various sectors under the radar, including agriculture, displacing more context-based or ecological approaches such as agroecology or organic farming. As technologies, they have been incorporated into a dominant discourse that promises a bright technological future; as such their justification is unquestioned. The discussion in the following chapters turns to formalist and structuralist ideas to develop a critique of the unproblematic acquiescence to a future defined by technologies of genetic manipulation, supported by neo-liberal policies. Formalist or structuralist ideas in biology and in the humanities have been rejected or marginalised by current postmodern trends. In particular, literary formalism developed in Russia and the United States in the early- to mid-twentieth century now seems dated and barely relevant to contemporary life. However, by turning to these formalist models, light could be shed on the issues surrounding genetic manipulation, particularly its assumptions of historical causation and inevitable progress. In biology, the blurring of the boundary between life and the inorganic, driven by the DNA revolution of the 1950s and 60s has led to

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

3

an unquestioned and unchallenged drive towards a bioengineered future. Formalist trends in both literary and biological studies are unique in their resistance to succumbing to the apparent inevitability of such ‘engines of progress’.

Dreams of Genetic Manipulation In Wes Jackson’s account, the approach of genetic technologies to agriculture is to address a specific problem with a specific solution while everything else is held constant. Molecular biology has a role to play in agriculture, Jackson suggests, but only in the context of the whole system. In industrial agriculture, there is a tendency to decontextualise agriculture from its biological system. This could be seen in the Green Revolution, which produced high-yielding varieties that had high requirements for external inputs, including fertilisers and pesticides. That this favours the interests of large chemical companies is no coincidence in the view of some commentators. It also favours larger landowners. The agrarian reformer, Wolf Ladejinsky, wrote in a report to the World Bank: “The Green Revolution is mainly confined to land with an assured water supply, larger holdings, and farmers with resources (both owned and borrowed) sufficiently large to take care of all other inputs” (Ladejinsky 1977). Meanwhile, soil resources are destroyed by an agro-technological approach to soil nutrition—methods to restore organic content, and even enhance it, with its associated microbial biota, have been known for centuries, but these too are side-lined by the dominant techno-centred discourse. A techno-biological approach is taken further as biology is harnessed by companies to cope with mountains of environmental waste generated by the same companies (Cooper 2008). Relevant to this is the first approval of a patent for a genetically engineered organism (and therefore the protection of propriety rights on a life form). This patent was approved by a US Supreme Court ruling for a bacterium modified genetically to degrade oil and therefore clean up oil spills. The controversial case Diamond vs Chakrabarty reached the US Supreme Court in 1980 and was widely reported at the time. According to Melinda Cooper, the movement of biology into waste management is generally driven by the very companies that generate waste—she refers to these happy arrangements as neoliberal utopias. Rather than an overtly authoritarian government, the process is promoted by business interests and, as Cooper (2008) outlined, a complete reformulation of the role of the biological sciences into the

4 

P. MCMAHON

handmaiden of neoliberal policies. But in the end a technological future, a biotechnological world that could concentrate power in the hands of a few is a dream held to by both capitalist democracies and autocratic states. Jackson raises the utopianesque dream of crops that can fix atmospheric nitrogen (N2) into soil-based (ammonium and nitrate) forms (the only forms of nitrogen that can be used by plants). An engineering possibility could be envisaged that taps into the abundance of nitrogen in the atmosphere (which is  over 70% nitrogen by volume). This could redress the perpetual shortage of soil nitrogen in agriculture (a nutrient that plants and we ultimately depend upon). If the genes (only present in legumes, such as peas and clover) that enable symbiotic associations with nitrogen-­ fixing bacteria Rhizobium could be inserted by genetic engineering into non-leguminous plants, such as wheat, a symbiotic relationship could be facilitated with the nitrogen-fixing Rhizobium. Problems of low soil nitrogen in broad-acre crop production could therefore be solved. This was an early claim of genetic engineers. As Jackson put it: “We were told then and we are told more now that this new biology …will make temperate cereal crops full blown nitrogen fixers when we stitch in legume genes to accommodate the nodule forming nitrogen-fixing bacteria of the soil” (Jackson 1984, 120). Currently, research to this end continues, in addition to attempts to incorporate bacterial genes coding for nitrogen-reducing catalysts directly into cereal genomes (Goyal et al. 2021). Potential problems are immediately raised by this scenario—abundance of N could lead to a process of eutrophication, observed when N fertilisers enter water bodies, leading to microbial population explosions, consumption of available oxygen and dead fish. Nutrient ratios which influence plant uptake would become unbalanced. A high soil concentration of N suppresses the uptake of other essential plant nutrients. High N/P and N/Mg ratios in the soil have been shown to suppress plant uptake of phosphorous (P) and magnesium (Mg), for example (Wessel 1971). High amounts of N promote vegetative growth, which would increase at the expense of grain or fruit production. Other issues are raised by genetic manipulation. Hans Jonas, commenting on visions of a future based on genetic manipulation, observes that bioengineering moves away from the experimental basis that has informed the biological sciences. It is no longer an experiment that can be scrapped or revised; the experiment merges with its real target. It is also irreversible, and cannot be recalled (unlike engineering failures that can be taken back to drawing table). If released into and successful in the wider environment

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

5

all changes are inherited in perpetuity (1974, 143). This already may be the case for recent releases of a species of mosquito (Aedes aegypti) genetically modified for reproductive failure. Aedes aegypti transmits a number of diseases, including dengue fever. These genetically modified (GM) mosquitoes (labelled OX5034) cannot successfully reproduce after mating, so that their release into the environment potentially leads to population decline, and therefore disease control. Hybrids formed between transgenics and wild populations of Aedes aegypti have been detected (Evans et al. 2019). Therefore, a portion of the genome of the GM mosquito could be incorporated into the wild population, although the company that developed the engineered mosquito claims the transgene eventually dies out when releases are ceased. Whether specialist insectivores, such as some bat species, will be impacted remains unclear. A further problem arises with the presence of antibiotics in the environment. The antibiotic, tetracycline enables the GM mosquito to survive; therefore, the presence of the antibiotic in water bodies could compromise the purpose of the release, which is to reduce the Aedes aegypti population. This is recognised by the US Environmental Protection Authority. Regarding releases of the GM mosquito in Florida it stipulates: Releases must not occur within 500 meters from the outer perimeter of 1) wastewater treatment facilities; 2) commercial citrus, apple, pear, nectarine and peach crops; and 3) commercial cattle, poultry, and pig livestock facilities. The 500-meter distance creates a conservative buffer zone between OX5034 release points and potential environmental tetracycline sources. (EPA release regulations reported in ‘Updates on GMO mosquito release’, August 8, 2023, Mercola.com)

An additional issue with genetic manipulation strategies is the claim that transgenic crops will lead to increased yields. These claims have not been realised (Gurian-Sherman 2009). Crops, including maize, have been genetically manipulated for herbicide resistance and for the internal production of compounds with insecticidal properties. The transgene conferring herbicide resistance to crops allows greater herbicide loads onto the environment. Attempts to increase yield by genetic engineering have been less successful. Most yield increases can be attributed to quantitative breeding methods. Additionally, environmental factors play a major role in yield but genetic manipulation attempts to bypass the agrosystem.

6 

P. MCMAHON

Selecting varieties suited to the actual local conditions would meet with greater success. An ecological approach to agriculture would view a crop plant as the property of nature and “a relative of wild things, [so] then we acknowledge that most of its evolution occurred in an ecological context, in a nature that was of a design not of our own making” (Jackson 1984). Genetically determined attributes in the wild relatives of crops and neglected crop cultivars are being lost through a process of genetic erosion. Crop cultivars have been selected for millennia by farmers responding to particular environmental conditions. Industrial agriculture replaces these cultivars with a few better ones which depend on intensive care (in other words, are insulated to an extent from environmental conditions). As traditional crop cultivars disappear, so do their particular genomes, some of which may have genes or suites of genes that express resistance to drought, pathogens and other environmental stresses. A tendency of ‘forgetting’ can be discerned in this agricultural approach: “[W]e are part of an age which may be said to be living on the accumulated capital—cultural and biological—of a million years of hardship, death, effort and invention … human society may behave irresponsibly for a time and forget the ties that bind it to the world” (Shepard 1967). Gene Technologies Have Potential Benefits Transformation technologies, such as genetic manipulation (engineering or modification),  gene editing, RNA interference, transplants, stem cell research, synthetic biology and others are now implicit to the discipline of biology. Modern biotechnologies, including genetic technologies, are useful and invaluable and the following chapters do not aim to criticise these promising technologies in themselves. Genetic modifications that are inherited via germ line (reproductive) cells or by asexual reproduction, including cell fission, should be distinguished from those that target body cells (with no effect on reproductive cells), such as gene therapy or prosthetic bioengineering. In the case of sexually reproducing organisms, it is important to differentiate somatic from germline methods of manipulation. The former methods modify non-reproductive (body) cells, and the latter target reproductive or germline cells. For example, somatic gene therapy has been trialled to mitigate genetic diseases, such as cystic fibrosis and haemophilia, by altering DNA sequences in the tissues of an individual afflicted with the condition. Only the body cells are altered. Here genetic

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

7

technologies show great promise. Nevertheless, the distinction (i.e. of soma and germline) may not be absolute, as pointed out by the evolutionary developmental biologists, Eva Jablonka and Marion Lamb (1995, 43): Sometimes somatic cells can ‘dedifferentiate’ and produce both stem cells and germ cells. This mode of development is commonly associated with asexual propagation. It is characteristic of all plants, all fungi and all multi-­ cellular protists, except the Volvacales. At least six animal phyla also show this kind of germ-cell formation.

However, generally the preeminence of the DNA code in the transmission of heritable characteristics is recognised by most biologists. This also applies to the propagation of plants by asexual means, such as tissue culture and grafting, as they result in the production of reproductive structures with an equivalent genetic composition to the somatic cells. The issue raised in this discussion is the implicit assumption that what is good for body cells (the soma) is even better if applied to reproductive cells—for then the changes will be permanent, passed on to following generations. With regard to medicine, major ethical considerations are raised—as Jurgen Habermas has pointed out such technologies put us on a slippery slope that ends up with designer babies and non-consensual changes to the genetic makeup of an individual. In agricultural and ecological applications of genetic technologies there is no such debate, or if there is it is quickly put down to a lack of understanding, disinformation or romantic ideals that have no place in efforts to address the agricultural or ecological problems that face society. Anything goes now in the sphere of (post)modern applications of biology in life forms other than humans. The organism and its relations are dismissed in the drive to a techno-biological future and the explicit aim is to alter organisms genetically at the germline level. It is therefore important to highlight alternative ways of looking at the organism, ways that question the value of the transformation of reproductive lines, while also finding value in other ways of applying genetic technologies. Somatic gene therapy could be an invaluable application in medicine. In agriculture, marker-assisted selection (MAS) uses molecular technology to assist quantitative breeding methods that take the organism (and its context) into account. This needs to be contrasted with direct genetic manipulation in agriculture, euphemistically called ‘molecular breeding’ (see Chap. 10).

8 

P. MCMAHON

Focus Is Placed on Practical Risks Critiques that question the premises behind genetic manipulation have been far and few between. One reason for this is postmodernist developments have moved away from the notion of developing critiques from an ‘outside’ vantage point so that “the external place for reflection and critique tends to disappear” (Lash 2002, 193). Humans become participants in interacting networks. As Bruno Latour put it, humans join the world as participatory ‘objects’ or actants: “Far from being “‘lowered down’, objectified humans will instead be elevated to the level of ants, chimps, chips, and particles!” (Latour 2005, 255). Laboratory life is now “externalised” into the “world out there” (Lash 2002, 192) and knowledge arises through new combinations. Taking a strongly empirical approach, Actor-Network Theory (ANT) postulates non-human actors including technological objects, enter into heterogeneous networks, each actor partially defining the other. The genetically modified organism (GMO) is a quasi-object, derived from networks of natural organisms, biotechnologies, social aspirations, political factors and business models. Another reason for the lack of questions regarding the premises of genetic manipulation technologies is that they embody and define the current and dominant discourse of biology. Direct manipulation is within, and part of, this discourse—and in agriculture is often portrayed merely as an advancement in breeding. Therefore, potential risks, drawbacks and other features of genetic technologies are assessed from within the paradigm, as they are  pored over by various committees. These evaluations address concerns over risks, compared to projected benefits and profitable outcomes, but the technology itself is an unquestioned part of the prevailing discourse. It is already considered to be another feather in the cap of humanism. Even where more voracious critiques are levelled against genetic technologies by environmental activists, for example, or informed citizens who raise food quality and safety issues, concerns are generally grouped around risks and the dangers associated with genetic manipulation, just as with any technology that poses environmental or health risks. More fundamental criticisms are dismissed as romantic, uninformed and even reactionary (Ferry 1995). An insightful essay by Anne Chapman (2005) attempted to develop a philosophical response to genetic engineering, based on the concept of nature as ‘other’. But generally, the discourse underpinning the technology is overlooked, with greater attention paid to its potential consequences.

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

9

As such, within biological and medical circles, critiques levelled at genetic manipulation usually cite its unforeseen consequences and its risks. These objections are based on consequentialist and utilitarian premises. They present little in the way of a priori arguments based on the organism or ‘types’. Some religious criticisms raise creationist beliefs in ‘natural’ kinds that should not be violated by interspecific crossings. Naturally, a number of counter-arguments arise in response to these religious critiques since the biological sciences demonstrate that the delineation between species is blurry, and is ill-defined in some cases (when can a population be deemed a species, or vice versa?). Furthermore, at the molecular level the similarity (in fact, the homology) between all life forms is striking. “It’s all the same” a La Trobe molecular biology professor, who actively promoted genetic engineering in Australia, once remarked. And, in terms of molecular mechanisms, he was right. The DNA in a mountain ash, a wombat, a cycad, and a hydra are identical in their basic structure and function.

The Code of Life and Bioengineering Since DNA, an inert and stable molecule itself, writes the ‘code of life’, the possibilities of bioengineering, artificial life and synthetic biology have come to the foreground in modern biological enterprises. The premise behind this, developed by Francis Crick (1916–2004) among others, is that biology in the final instance is based on the laws of physics and chemistry, and the boundary between life and non-life is blurred, even non-­ existent. Crick believed that eventual acceptance of this ‘truth’ would lead the way to a future of beneficial bioengineering. Francis Crick stated that with the advent of molecular methods, all biology would be explainable in terms of physics and chemistry. This ultimately, Crick thought, “would open up the world of bioengineering” (Petersen 2023, 176). Crick’s endorsement of physico-chemical explanations in his ‘Of molecules and men’ demonstrates his belief in the lack of a hard border between the living and non-living—viruses, for example, are essentially ordered chemical factories. Life’s origin was most likely a chemical soup in which proto-forms of the molecules of life, nucleic acids and proteins, were spontaneously produced. Evolutionary ideas of natural selection also came down in the end to chemistry, he proposed, in mutations or alterations in the DNA code. Crick believed that bioengineering would (1) finally put to rest the notion of a life/non-life border, (2) explain the origin of life as a chemical

10 

P. MCMAHON

mix and (3) even explain consciousness in molecular terms. In some ways there is a blurry boundary between life and the inorganic world. D’Arcy Thompson showed this clearly: the light bones of the bird, the length/ breadth ratio of supporting limbs, the current deployed by electric eels, or the sonic perception of bats, as well as the laws of chemistry being equally applicable to the biochemistry of the cell, all show how life is in a continuum with the inorganic. It was Crick’s belief and hope that one day living processes could be accounted for solely in terms of chemistry, and ultimately, physics. Yet the life/non-life border refuses to be deconstructed— to Thompson and other structuralists life has a non-reducible form, leading to their relegation to the vitalist camp. Crick’s predictions have come to fruition with the rise of genetic manipulation technologies and more recent developments, such as synthetic biology and artificial life. The eventual aim of synthetic biology is to construct living cells from organic molecules, although a first step in achieving this is to combine living cytoplasm with foreign genomes. Crick’s insistence that organic life can only be understood through an understanding of its inorganic base, provides the philosophical underpinning, the carte blanche, for these technologies. While biosystems are extremely complex, higher levels of organisation are explained in terms of molecular interactions and their emergent properties. By this narrative, in the end all life can be explained by applying basic physical and chemical principles. The code of life is common to organisms with very different structural arrangements. Eukaryotes (complex organisms, including humans) and the simpler prokaryotes (bacteria and archaea) are based on the same coding system. It is clear that, at the genetic level, it is difficult to defend the concept of clearly delineated kinds or species. We share most of our DNA with our closest relatives, the non-human primates, and much of our DNA and even functional genes overlaps with that of other eukaryotes, including fungi, trees and algae. Despite this, mechanisms of isolation (whether genetic or reproductive) ensure the separation of populations and (according to neo-Darwinist theory) their divergence into distinct species. According to Ernst Mayr, reproductive isolation (the separation of populations so that interbreeding is prevented) is the primary mechanism in speciation. Both allopatric isolation due to geographical barriers and sympatric mechanisms of speciation (such as niche separation) may have contributed to the diverse and heterogeneous biosphere essential to the function of our planet (Maynard-Smith 1966). This is the best explanation

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

11

for closely related species that emerge due to a lack of exchange of genes (referred to as sister species) but questions remain with regard to the emergence of very different structures (or body plans). Other explanations of macroevolution have been proposed (Chaps. 7 and 8). Structural Relations and Vitalism The notion of the organism as having structural unity was raised by Immanuel Kant, who differentiated it from the functional unity of the machine. Our current approach to biology remains fundamentally Cartesian—treating the organism as a functional machine, one that is complicated and challenging to understand to be sure, but something that in the end we can unravel, even replicate, as a mechanical or computer information system. A structuralist approach, on the contrary, emphasises constitutive relations; this has been applied in biology by theoretical biologists, questioning the current status of the gene as the primary functional unit of life. Erik Petersen maintains that according to Francis Crick vitalists were “were anyone who thought he was wrong about the ease of reduction or who disagreed with the goal of engineering life” (Petersen 2023, 183). Crick’s attack on vitalism was closely related to his “vision of a techno-­ logically mediated destiny” (ibid., 175), a future based on bioengineering. Crick saw a future of synthesised organisms, engineered for human good. In this sense, he was a self-acknowledged humanist along the lines of Francis Bacon in the late Renaissance. The future would replace biological explanations with exclusively physico-chemical ones, meaning that “biological functions, including cognition, would be replaced by engineered, synthetic cells and computers” (ibid., 175). As Petersen recounted, the famous division between the sciences and humanities (the ‘two cultures’) lamented by C.P.  Snow was for Crick a triumph of materialist science, which would eventually explain all phenomena, including consciousness itself. In this regard, his stance was consistent with the physicalism of the Vienna logical positivists. It also has something in common with evolutionary psychology, which adopting an adaptationist approach, links human behaviour to a material process of natural selection and gene action. Some theoretical biologists, notably Robert Rosen, have objected to characterising life as reducible to physical and chemical principles. Using set theory they demonstrate that life operates on internally derived principles, although also working as an open thermodynamic system that

12 

P. MCMAHON

sources materials from the environment. Rather than vitalism, the old idea of form (held by early biologists) is returned to—although endorsing form in itself has been accused of being vitalistic (both Hans Driesch and d’Arcy Thompson were accused of vitalism, perhaps justified in the case of the later Driesch due to his promotion of entelechy). To Rosen and relational biologists, the principle of life is based on its organisation (Chap. 5). This organisation is not only mechanical but also relational (or constitutive) since systems within the cell are constituted by other systems without any need for an external input. For this reason, relational biologists critique the notion of artificial life (AL) since no computer (program) has been developed that does not need informational input from an external source (Cardenas et al. 2010).

Structuralism Across Disciplines Whilst structuralist ideas in biology have tended to be marginalised, structuralism has been an important driver of linguistic studies, following the linguistic turn of the early twentieth century. The two disciplines (biology and linguistics) operate in totally different spheres (one physical and the other symbolic); nevertheless, Ernst Cassirer (1945) suggested common themes may be found between biological and linguistic structuralism. In his ‘Structuralism in Modern Linguistics’, a paper presented to the Linguistic Circle of New York in 1945 just before his death, Cassirer strongly criticised the characterisation of language as a ‘thing’ that could be subjected to the same kinds of analysis used by chemists and physicists, or as an organism. August Schleicher, for example, had described language as a living organism, with a birth, life and final demise (Schleicher 1863 Die Darwinische Theorie und die Sprachwissenschaft). Cassirer reminded his audience that language is strictly symbolic and requires analysis of meaning, not of its physical properties, for example, phonetics. He quoted the Russian structuralist, Nikolai Trubetzkoy, one of the founders of the Prague linguistics school, who emphasised the division between phonetics and phonology, the former belonging to the province of physical sounds, and the latter in the incorporeal realm of signification. Nevertheless, common themes arise in biology and linguistics, says Cassirer: Biologists and linguists are often engaged in the same battle against a common adversary, a battle that may be described by the slogan: structuralism versus mechanism; morphologism against materialism.

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

13

The linguist Wilhelm von Humboldt, a member of the Weimer group, applied Johann von Goethe’s principles to his own ideas. The principle of a durable morphology in plants was also subject to change and diversity, expressed by Goethe’s lines in ‘The Metamorphosis of Plants’: Like unto each the form, yet none alike; And so the choir hints a secret law.

Humboldt applied this to his concept of linguist types. As Cassirer put it in his paper: “Goethe sought for this hidden law in the natural world; Humboldt tried to discover it in the cultural world, the world of human speech.” Those adhering to the discourse of positivism that dominated the nineteenth century were suspicious of such views, said Cassirer. This could be said equally of some of the critical stances towards structuralist biology in current times. But in linguistics, structuralism had made a comeback by the mid-twentieth century. And to Cassirer, structuralism “is no isolated phenomenon” but rather “a general tendency in thought” (ibid.). Alternative ideas in biology, particularly in relation to evolution, have persisted for a long time, albeit on the margins of biological thought. E.S. Russell (1916) stated: It is questionable indeed whether the theory of natural selection is properly applicable to the problem of form. It was invented to account for the evolution of specific differences and of ecological adaptations, it was not primarily intended as an explanation of the more wonderful and more mysterious facts of the convenance des parties and the interaction of structure and function. (Russell 1916, 232)

And Charles Darwin himself did not seek an explanation of homological form in adaptation: Nothing can be more hopeless than to attempt to explain the similarity of pattern in members of the same class, by utility or by the doctrine of final causes. (Darwin 1859, 434)

Now in an ‘extended synthesis’ Darwinian theory grapples with gene homology as the degree of conservation of regulatory and structural proteins is shown to be unprecedented (dispelling the earlier predictions by evolutionary theorists based on adaptationism). Such homology is thought

14 

P. MCMAHON

to reflect not only resemblance, but natural units of form (Wagner 1999). Genes are themselves ancient structures and differential expression (the use of the same gene template to produce different proteins) requires many proteins (50–100), including chromatin-modifying proteins, transcription factors and their co-factors. Particularly well known are the Homeobox (Hox) genes, a family of transcription factors: these regulators act to position cells along an anterior-posterior axis in development. They are highly conserved between different phyla (from insects to mammals), lending credibility to the position of Etienne Geoffroy Saint-Hilaire (1772–1844) that common structures link living things, at least in eukaryotes (i.e. organisms more complex than bacteria) (Carrol et  al. 2001, 213). In addition, developmental plasticity is thought to be an important mechanism in evolution, as phenotypic variation becomes fixed and selected. Conrad Waddington showed that environmental conditions could trigger a change in eye colour in Drosophila fruit flies, which was then inherited, that is, fixed by selection, a process known as genetic assimilation (Chap. 8). And Mary Jane West-Eberhard has proposed that “genes are probably more often followers, than leaders in evolutionary change” (West-Eberhard 2005). Structuralist biology has taken up general principles that were applied across various disciplines culminating in the ‘high structuralism’ that dominated the humanities in the 1960s (Chap. 6). It is also informed by French ideas debated within the new science of biology in the early nineteenth century, and returned to at the end of the century in the reaction to social Darwinism, realism and mechanism. Structuralists reject atomistic explanations of function, and in biology are critical of the premises underlying genetic manipulation, which suggest each gene is linked to a function that can be transferred from one species to another. Taking up Ernst Cassirer’s proposal that comparisons can be made between structuralist theory in biology and the humanities, the aim of the discussion in this book is to examine some of the formalist premises of alternative biological accounts, with comparisons to similar themes in literary theory. Formalism Returns with the Modernist Movement At the end of the nineteenth century, a strong reaction against Victorian positivism and explanations in terms of historical origins was expressed with the rise of modernist art. New art forms such as imagism in poetry and cubism in painting broke onto the scene. Sciences were also affected

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

15

and biology, during what has been called the ‘eclipse of Darwinism’, returned to some of the earlier rationalist ideas that had been overtaken by Darwinian thought and adaptationism. Darwin’s theory was based on a gradualist and cumulative account of evolutionary change as organisms adapted to their conditions of existence. The theory morphed into more extreme expressions of adaptationism, which was presented as an explanation for all biodiversity. Critiques of the Darwinian adaptationist program in biology emerged in the early modernist era (a backlash to the excesses of Victorian mechanism). In particular, scepticism was directed towards the theory that speciation could result from the gradual accumulation of units of variation. Henri Bergson wrote Creative Evolution (Bergson  1911). Organicist ideas that rejected the divisibility of the organism were put forward, for example, in ‘On Growth and Form’ by D’Arcy Thompson (1917). During this period, a Swiss teacher and linguist Ferdinand de Saussure (1857–1913) developed his systems theory of language in response to nineteenth-century linguistics, rejecting its historicity and search for origins. He was critical of the focus on the diachronic (changing) nature of language, wishing to identify a core system that could be applied across individual languages and histories. Other revisions of the historical approach occurred for similar reasons: Edmund Husserl developed a theory of logic that did not turn to history or origins for explanation. Sigmund Freud believed the unconscious could finally be accounted for in scientific terms. A broad reaction against the prevailing belief in a ‘history’ of inevitable ‘progress’ under Enlightenment ideals, characterised the modernist movement. The striking thing about form in some modernist art and literature is its simplicity. Imagist modernist poetry explores form with short refrains and contrasting images which are united in meaning. The frequently cited example is Ezra Pound’s image of a railway station platform in one of his poems: The apparition of these faces in the crowd: Petals on a wet, black bough.

Modernist art explored form in elegant simple designs or outlines or the brutal fractured cubism of Picasso and others. Later, Ernest Hemingway developed a brevity in form that was spare and direct. Such directness spurned the verbose, flowery and often mendacious language exemplified

16 

P. MCMAHON

in earlier literature or texts. New Criticism analysed texts from within, solely based on their form and associated meaning, rejecting any contextual, social or historical accounts (Belsey 2002). The particular form of individual texts provided their meaning. Russian literary theorists, such as Vladimir Propp, took a different approach, finding universal relations in fairy tales, so that figuratively without ‘the princess and evil stepmother’ there could be no ‘hero’. Each protagonist is defined by its place in relation to the others in the tale. A significant feature of this movement was a rejection of explanations based on history and the inevitable train of progress—modernism was imbued with scepticism in an era both preparing for and recovering from war and a feeling of hopelessness, along with the rise of impersonal industrialisation and bureaucracy. Ezra Pound, for example, declared it was time to ‘make things new’. Again, simplicity and directness were the order of the day. Greater interest was displayed in the process of perception and the ordering of form by the subconscious. Now inner life had come to the fore, and the subconscious was a key term (rather than self-consciousness). The individual was directed to an extent by inner forces, illustrated by T.S. Eliot’s ‘patient etherised upon the table’. Modernism’s predecessor in spirit was perhaps Romanticism. Both displayed a reaction against earlier discourses, the materialism of the Age of Reason, emerging industrialisation, increasing wealth for the few and the great machines of war. However, Romantics attached significance to nature or an idyllic past, both representing innocence and purity. Romantic introversion was melancholic, nostalgic, even escapist. It was also preoccupied with natural relations and the organic body. Modernist artists showed no attachment to Nature or History, and were more interested in exposure of Enlightenment ideals and the resulting militarism and exploitative industry as futile and even dangerous. They aimed to expose the inner bones of existence, spurning frills and excess in language. Structuralism took as its starting point the linguistic turn of the early twentieth century, and in particular the systemic model of language proposed by Ferdinand de Saussure. Language, as a system or langue, was at the base of its use (speech or parole). We only have direct access to language as uttered, but Saussure connected speech to an underlying structure, that he termed ‘langue’. The revolutionary possibilities of Saussure’s linguistic theory were adopted by a range of structuralist discourses (Chap. 6). Structuralist theory is concerned with the whole system, which is self-­ sufficient in that internal relations in the system (rather than relations with

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

17

an outer ‘real’ world) define the meaning produced in the system. Language, the pre-eminent structure, is based on social relations that precede the individual subject (Matthews 1996). Meaning is not personal but belongs to Other (or language) and is therefore shared; in that sense meaning belongs to other people (Belsey 2002, 43). Formalist and structuralist ideas affected various disciplines, including literature and biology, rejecting the notion that an agent, such as the ‘author's’ intention or the ‘gene’, could explain meaning or direct function. An interest returned to the biological form that had concerned studies a century previously, along with a critique of historico-adaptive explanations of life. The modernist turn brought with it a resurrection of previous (pre-Darwinian) biological theories, particularly those debated in France in the early nineteenth century. A greater focus was placed on the internal form of the organism, along with a return to vitalist theories. This was in part a reaction against the materialist approaches to the biological sciences that were predominant in the late nineteenth century. Biological structuralism shares features with the linguistic ideas that emerged at about the same time. In common with structuralist linguistics, biology of the late nineteenth century rejected historicist explanations, with a renewed interest in permanent (or long-lasting) form. Indivisible Form Modernist thought and its effect on biology was influenced by the idea of formal relations, whether these were directed by function or by universal ‘laws of form’. The merging of form and function can be seen in modernist architecture. Form was brought to the level of function—so that the proportions (or relations) in an oblong room were set as most suited to its function. Extra frills were not acceptable. In response to this, postmodern architecture embodies a reaction against pure functionalism, and tries to reintroduce and eclectic mix of period themes and decor. Simplicity in biological form has been raised by Brian Goodwin, who commented that while the mathematics describing formal relationships is extremely complex, forms are themselves elegantly simple (Goodwin 1990). Other rationalist views also sought simple explanations based on principles of connection and balance. In both modernist and organic relations, function reveals a simplicity of formal relations. Modernist art explicitly attempted to express sparsity of basic forms; a glimpse behind the external veneer.

18 

P. MCMAHON

Structuralist biology in the early modernist period displayed a renewed interest in (ahistorical) laws of form in organic nature. D’Arcy Thompson’s ‘On Growth on Form’ published in 1917 is, according to one commentator, becoming increasingly relevant to current debates within biophilosophy (Karsenti 2008). D’Arcy Thompson examined living form as an engineering problem, for example, the truss-like structures in birds’ hollow bones, that ensure lightness combined with strength. Thompson was particularly interested in the physical constraints that limited development of organisms: he applied mathematical ratios such as the rate of increase (growth) of linear or cubic dimensions—the latter explains why a flamingo does not have legs that are 10 feet long, when compared to a stilt: rather than proportions (such as height of the torso compared to legs in birds) being expressed in linear terms, nature in this case reflects the rate of increase of cubic volumes. Herbert E. Huntley (1888–1976) discussing Thompson in his ‘The Divine Proportion’ mentions the spiral structures of pine cones, pineapples and leaf arrangements. Such phyllotaxis is based on the Fibonacci sequence—indicating branching in life forms has a mathematical basis (Huntley 1970) or “mathematical field-like structures that seem to recapitulate biological patterns” (Levin 2012). D’Arcy Thompson was interested in length and strength relationships—so elephants are ungainly if compared to a gazelle; they need to have a higher width-length ratio in their legs to support the greater weight. The size of the elephant approaches the limit of possible dimensions for terrestrial animals, Thompson suggested, so long as they are not partially submerged in water (thought to explain the massive size achieved by some extinct reptiles) or other possible factors, such as hollow (lighter) bones characteristic of some dinosaurs. His most famous chapter was concerned with mathematical transformations, which he claimed could be traced to the relations plotted on a Cartesian coordinate grid (Fig. 1.1). Due to his interest in the mathematically consistent patterns apparent in living forms, and a wholesale rejection that these evolve ‘part by part’ (the position of Darwinian adaptationists), he has been referred to as a vitalist. However, formalist positions are often lumped together with vitalism, as discussed further below. Form Is Conflated with Vitalism Ideas of form in living systems, including organicism, are often grouped with vitalism. Charles Wolfe (2011) differentiates functional and

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

19

Fig. 1.1  An example of D’Arcy Thompson’s proposal of homology based on transformation: the sunfish on the right (b) could be considered a deformation of Diodon, the porcupine fish, left (a) as homologous points on each fish species have similar coordinates on the deformed grid. (Creative Commons, Wikimedia)

substantival (or metaphysical) vitalism: the latter endorses a non-material life force, while the former states that life (and how it functions) cannot be deduced from inorganic principles (Moir 2023). Examples of metaphysical vitalism are putative extra-material forces or flows, such as Henri Bergson’s elan vital and Hans Driesch’s entelechy. In his famous experimental work on the development of sea urchins, Driesch separated cells of the sea urchin at the earliest stage of development when the fertilised egg (zygote) had divided, and left them overnight. He assumed based on the research of Wilhelm Roux that each cell would be predetermined (preformed) to develop into only a part of the sea urchin larva: “Instead, the next morning I found in their respective dishes typical, actively swimming blastulae of half size” (Driesch 1892 [1964, 46]; quoted by Fagan and Maienschein 2022). If the cells produced when the zygote first divided were separated, Driesch found that each cell could develop into a whole organism. He later proposed this lent

20 

P. MCMAHON

support to the old Aristotelian concept of entelechy, an actualisation of potentiality, that accounts for living function. Driesch’s entelechy is an instance of metaphysical vitalism according to Wolfe. It has been criticised by structuralist biologists, referring to it as “a non-material, non-energetic and non-spatial agent …which, in the last analysis, seems to be effectively an idealist ‘central directing agency’ and about which little of a positive nature can be said” (Webster and Goodwin 1982, 40). However, the obvious conclusion of Hans Driesch’s experiment was that the embryonic cells are totipotent (each being able to undergo material transformations resulting in a whole organism). His work had an impact on the perennial epigenesis/preformation debate (whether an embryo undergoes material transformations, or is simply a smaller replica of the adult). Driesch’s embryological research lent support to a theory of development based on a transformative process in the constituent materials of the organism, the notion of epigenesis. Proposing that material could change itself conflicted with the orthodox view of preformation (that a small replica of the adult existed from the start in embryo development). Marjorie Grene and David Depew point out that among early concepts of embryology the Church favoured a preformation model, which proposed that a mini-version of the adult was enclosed in the embryo: “Somehow, all the plants and animals that ever would be were contained in little in the first of their kind when God created them” (Grene and Depew 2004, 83). Such an explanation fitted in with living forms as divine creations. Epigenesis, on the other hand, proposed that material could undergo change, or that the embryo developed from something quite different, a view that the theology of the time found threatening to its position, posing the question “Are we to have spontaneous generation without God’s creative power?” (Grene and Depew 2004, 86). Thus the clerics promoted the reality of species, based on eternal form, while ideas such as transmutation or epigenesis threatened this assumption of stability, so were strongly refuted by church authorities. Among the prominent French scientists who put biology on the map in the early nineteenth century (Chap. 2), Georges Cuvier, supported preformation, while his colleagues, Jean-­ Baptiste Lamarck and Geoffroy Saint Hilaire, among others, took a stance closer to the materialism of the deist (and atheist) beliefs that characterised late Enlightenment thought. Driesch, more than many, finally settled this debate on the side of epigenesis, although his own proposal of an entelechy, or extra-material principle, has been largely rejected.

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

21

In contrast to metaphysical vitalism, organicism, equated with functional vitalism, is nothing more than biological organisation, an organisation unique to life that separates the organic from the inorganic (Wolfe 2011). In this sense, it is reminiscent of Georges Cuvier’s findings in his anatomical studies of an integrated organisation, which sets a clear boundary between life and the inorganic environment (Chap. 2). It is also endorsed by relational biology proponents, who claim that biology is underpinned by unique principles, not reducible to physics or chemistry (Chap. 5). Cassirer’s Critique of Both Vitalism and Reductionism Ernst Cassirer played an important role in critiquing vitalist theories, without subscribing to the reductionist account presented by scientists of the Vienna Circle of logical positivism, with whom he was closely associated. However, he was also influenced by the ideas of Jakob von Uexkull, a German-Estonian biologist who developed the concept of Umwelt (Stjernfelt 2011)—see Chap. 9. Cassirer favoured the structural concept of the whole, since it avoids teleology (i.e. that life moves towards a pre-set end) but supports ‘purposeless purposiveness’ (which Immanuel Kant differentiated from teleology). In other words, ‘form’ replaces the ‘final cause’. Agreeing with von Uexkull, Cassirer supported the study of physics as an investigation into causality, but also suggested that not all scientific phenomena are causal (Chirimuuta 2023, 97). To Cassirer, the reductionism of the Vienna scientists was forgetful of morphology, of the structure of whole organisms. Cassirer took up Johann von Goethe’s and von Uexkull’s emphasis on morphology as the arena for study of life (the organism), rather than its constitutive components. In an effort to separate humanities and science (which Vienna scientists aimed to merge into one broad discipline based on the principle of physicalism) he associated the humanities with form, placing biology alongside the humanities. The aim was a science of form, a form that is not causal, suggesting that physics and biology have different premises. As Marjorie Grene points out, the deductive approach in physical sciences was now replaced by an inductive method in the study of living things, since the organism is already present and fully formed (Petersen 2023). However, von Uexkull, who inspired Cassirer, presented form in a particular light, something akin to mathematics. He compared living form to

22 

P. MCMAHON

geometry, that is, an immaterial form (similar in a way to d’Arcy Thompson’s idea of form as a set of Cartesian coordinates—see above). This set von Uexkull and Cassirer apart from the Vienna scientists who viewed this as vitalism. In response, Cassirer brought up lessons from physics such as the electromagnetic ‘field’ and the electron, neither identifiable as physical elements nor as ‘parts’. His criticism of Vienna reductionism was similar to Conrad Waddington’s review of Francis Crick’s publication, ‘Of molecules and men’ where he alluded to the quark (what is it, in fact? Waddington asked—nobody can define it), highlighting the difficulties of relying on reductionism in physics (Petersen 2023). Rudolf Carnap (1891–1970) and Ernest Nagel (1901–1985) argued strongly against vitalism and organicism but suggested that biology had not reached the stage of reducibility (seeing it as a work in progress). The idea that there could be unique biological laws was portrayed as mystical nonsense: “Thus Carnap rules out vitalism and organicism on the same grounds—both of these doctrines posit biological concepts, e.g. entelechy for vitalism, the concept of the ‘whole’ and ‘organism’ for organicism, which evade non-metaphysical construal” (Chirimuuta 2023, 93). They sought the unity of science based on physicalism (the unitary science sought by the Vienna positivists), creating tension with those who believed in autonomy of biology and life: “Whereas mechanism is consistent with physicalism and the unity of science, proponents of vitalism and organicism assert that biology is in some sense autonomous from the physical sciences—that it relies on its own laws or concepts, lacking correspondence to ones stateable in the physical language” (ibid., 93). Ernst Cassirer subscribed to organicism (rather than a metaphysical vitalism), rejecting both the notion of a non-material force (the elan vital or entelechy) and the physicalism of the Vienna Circle. With regard to the latter, he claimed such ideas lose the ‘phenomenon’, for example, a painting is not just flecks of paint on a canvas, but holds meaning. Vitalism is a label most biologists go to great lengths to avoid, yet in doing so can only turn to reductionist ideas such as those put forward by the Vienna circle of logical positivism. This is because with the turn to hard reductionism in the biological sciences of the mid-late twentieth century, a vitalist was defined as anyone veering from reductionist criteria in attempting to define life. However, those who endorsed organicism, such as Ernst Cassirer, were themselves strong critics of the vitalist notion of an extra-material force, such as elan vital or entelechy. Cassirer recognised that Driesch (with his sea urchin experiments) had identified a problem

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

23

unique to biology, but did not accept his entelechy, even though Driesch himself tried to invoke its principles from physics, especially the notion of cause and effect. Although Cassirer and Martin Heidegger had many disagreements culminating in their famous meeting at Davos, Switzerland (Sullivan and Malkmus 2016), they concurred in disagreeing with Driesch’s attempt to base his arguments on principles of causality. Against Driesch, Cassirer reminded that Kant’s purposiveness in organic beings is a point of view, not an inherent power. He turned to Kant’s idea that the organic cannot be explained in terms of physical causality. Similarly, Martin Heidegger reminded that organs are not the same as parts of a machine and in the case of Driesch claimed he had merely handed over the problem to a causal factor, a force (entelechy). Thus the problem remains an open question—entelechy is only a way of closing the question. This force “explains nothing” stated Heidegger (1995, 262). Ernst Cassirer also criticised Henri Bergson’s idea of intuitive knowledge. Bergson proposed that a vital impulse (elan vital) flowed through life and that could not be apprehended intellectually, but only intuitively (Bergson 1911). Bergson placed immediate intuition as primary to intellectual activity (the latter consisting of rigid and lifeless concepts). Intuition was therefore associated with perception and intellectualism with conception. In that sense, Bergson’s ideas had strong similarities to those of Goethe. According to Chirimuuta (2023), Cassirer wanted to pull Goethe towards the neo-Kantian camp. Cassirer associated Bergson’s vitalism with mysticism, asserting that philosophy is “closed to immediacy” (ibid., 89) and endorsed a Kantian mediacy (through symbolism) instead. Cassirer developed Kantian ideas into his own symbology (with symbols as mediators of knowledge). Boundary of Life/Non-life While organicism, and also the vitalism of Driesch, puts emphasis on the boundary between life forms and their inorganic environment (found also in the thought of Georges Cuvier—Chap. 2), both panvitalist and extreme adaptationist models of organisms blur this distinction. On the vitalist side, Jean-Baptiste Lamarck proposed a form of vital materialism that applied to inorganic, as well as, organic bodies, a form of hylozoism. Hylozoism (or panvitalism) ascribes life to all matter. Lamarck believed in an inherent drive towards increased complexity. Lamarck based his progressive ideas on the rationalist philosophy of the eighteenth

24 

P. MCMAHON

century, viewing life as changing from the simple to the complex, with humans at the pinnacle of this chain; orthogenetic evolutionary ideas are based on a progressive trend to increased complexity to a certain end. A century later, Bergson also subscribed to orthogenesis, expressed in his ‘Creative Evolution’, which implies a directed evolution. And Simon Conway Morris has pointed to striking examples of convergent evolution (or the evolution of similar structures in unrelated taxa) that support the ideas of pre-set pathways of evolution. Therefore, to Simon Conway Morris, in opposition to proponents of an undirected evolution such as Stephen Jay Gould, the evolution of humans (or human-like organisms) was inevitable since evolutionary pathways are constrained by pre-existing laws or pathways. Ernst Haeckel also developed evolutionary ideas based on orthogenesis, or a progressive evolution. Haeckel’s hylozoism was based on a mechanistic view that life was an outcome of common general laws (unifying principles)—and a monism (Moir 2023, 244): “It was by erasing the distinction between life and non-life that Haeckel could argue that everything from crystals to human societies operated in accordance with ‘natural’, transhistorical evolutionary laws” (ibid., 246). Haeckel’s views are considered to be Lamarckian, but he was also labelled a social Darwinist, due to his ideas on the superiority of some human races, a forerunner of national socialist beliefs in superiority of particular Caucasian races. His hylozoism and monistic ideas inspired the establishment of the ‘monist league’; however, the monist league was explicitly anti-Nazi, accusing the regime of anti-scientific beliefs, it eventually disbanded in the face of Nazi coercion of institutions. The anti-vitalism emerging with the molecular biology of the mid-to-­ late twentieth century, promulgated by Francis Crick and others, led to a refusal to attribute unique properties to life. Here, again (just as in the vital materialism of Lamarck or Haeckel), no strict division was postulated between life and non-life. Panvitalism (hylozoism) proposes that life imbues all phenomena, blurring the boundary between life and its inorganic milieu. And molecular reductionism, similarly, refutes an autonomy in life that separates it from the inorganic. Scott Lash suggests vitalism has reemerged in postmodern models, which postulate a flow between the living and non-living and “splices of the natural and artefactual” (Lash 2002, 193). Gilles Deleuze engaged with Bergsonian vitalism to develop his own views on the ‘virtual’. The ‘virtual’ is counterposed to the ‘actual’—so that the germline (Weissman’s

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

25

immortal hereditary mechanism) becomes associated with the virtual (a non-divisible, constitutive force), and the body, organisms, cells or trees, with the actual. The virtual as suggested by Deleuze is the genetic flow of information that occurs between generations: “From virtual tendencies or potentialities certain beings are actualised. Genes, for example, actualise themselves in distinct bodies but also harness powers for further mutation and becoming beyond the body which expresses them” (Colebrook 2002, 96). Genes are the repository of information passed on from generation to generation. The organisms are the actual bodies resulting from this vital flow. In other words, flows of genetic information are granted a primary and formative place in relation to living cells and organisms. The organism or the body is therefore a temporary vehicle for the immortal ‘selfish’ gene. These ideas (paradoxically) have something in common with the reductionist materialist position held widely in molecular biology—namely that the boundary between life and non-life is essentially non-existent. The removal of boundaries, including those that define the organism, is the defining feature of postmodernism and its biological applications (Haraway 1991). Therefore, an ironic situation develops whereby the rejection of vitalism by the field of molecular biology is replaced by a new kind of vitalism in (post)modern biology, one that drives genetic manipulation: “[G]enetic coding seems to be almost an extension of the coding of media and message. If classical vitalism conceives of life as flow and in opposition to the structures that would contain and stop it, neo-vitalism would seem to have its roots in something like a media or information heuristic, thus there is talk today that ‘information is alive’” (Lash 2006). This new kind of vitalism is driven by business interests, but is also founded on the postmodernist deconstruction of boundaries, including the life/ non-life division.

References Belsey, Catherine. 2002. Critical practice. 2nd ed. New York: Routledge. Bergson, H. 1911. Creative evolution. London: MacMillan and co. Cardenas, M.L., J.C.  Letelier, C.  Gutierrez, A.  Cornish-Bowden, and J.  Soto-­ Andrade. 2010. Closure to efficient causation, computability and artificial life. Journal of Theoretical Biology 263 (1): 79–92. Carrol, Sean B., Jennifer K. Grenier, and Scott D. Weatherbee. 2001. From DNA to diversity: Molecular genetics and the evolution of animal design. Oxford: Blackwell Science Ltd. Cassirer, Ernst. 1945. Structuralism in modern linguistics. WORD 1: 99–120.

26 

P. MCMAHON

Chapman, Anne. 2005. Genetic engineering: The unnatural argument. Techne 9 (2 Winter): 81–93. Chirimuuta, M. 2023. The critical difference between holism and vitalism in Cassirer’s philosophy of science. In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T.  Wolfe, 85–106. Cham: Springer. Colebrook, Claire. 2002. Gilles Deleuze. New York: Routledge. Cooper, Melinda E. 2008. Life as surplus: Biotechnology and capitalism in the neoliberal era. Washington, DC: University of Washington Press. Darwin, Charles. 1859. On the origin of species. 1st ed. Harvard UP. Driesch, Hans, 1892 [1964]. “Entwicklungsmechanische Studien. I. Der Werth der beiden ersten Furchungszellen in der Echinodermentwicklung. Experimentelle Erzeugen von Theil- und Doppelbildung”, Zeitschrift für wissenschafliche Zoologie, 53: 160–178. Abridged translation as “The Potency of the First Two Cleavage Cells in Echinoderm Development. Experimental Production of Partial and Double Formations.” L. Mezger, M. Hamburger, V. Hamburger, and T.S. Hall (trans.), in Willier and Oppenheimer 1964: 38–50. Evans, B.R., P.  Kotsakiozi, A.L.  Costa-da-Silva, R.S.  Ioshino, L.  Garziera, M.C.  Pedrosa, A.  Malavasi, J.F.  Virginio, M.L.  Capurro, and J.R.  Powell. 2019. Transgenic Aedes aegypti mosquitoes transfer genes into a natural population. Scientific Reports 9 (1): 13047. https://doi.org/10.1038/s41598-­ 019-­49660-­6. Fagan, Melinda Bonnie, and Jane Maienschein. 2022. Theories of biological development. In The Stanford Encyclopedia of philosophy (summer 2022 edition), ed. N. Edward. Zalta. Ferry, Luc. 1995. The new ecological order. Trans. Carol Volk. Grene, Marjorie, and David J. Depew. 2004. The philosophy of biology: an episodic history. New York: Cambridge University Press. Goodwin, B.C. 1990. Structuralism in biology. Science Progress (1933–) 47: 227–244. Goyal, R.K., M.A.  Schmidt, and M.F.  Hynes. 2021. Molecular biology in the improvement of biological nitrogen fixation by rhizobia and extending the scope to cereals. Microorganisms 9 (1). https://doi.org/10.3390/microorganisms 9010125. Gurian-Sherman, Doug 2009. Failure to yield. Union of Concerned Scientists Report. Cambridge MA: UCS Publications. Habermas, Jurgen. 2003. The future of human nature. Polity Press. Haraway, Donna Jeanne. 1991. Simians, cyborgs, and women: The reinvention of nature. London: Free Association. Heidegger, Martin. 1995. The fundamental concepts of metaphysics. Trans. William McNeill and Nicholas Walker. Bloomington: Indiana University Press. Huntley, Herbert E. 1970. The divine proportion. New York: Dover Publications.

1  STRUCTURALISM, VITALISM, AND BIOENGINEERING 

27

Jablonka, Eva, and Marion Lamb. 1995. Epigenetic inheritance and evolution: The Lamarckian dimension. Oxford, New York: Oxford University Press. Jackson, Wes. 1984. A new threat to agriculture and a window of opportunity. The Ecologist 14 (3): 119–124. Jonas, Hans. 1974. Philosophical essays from ancient creed to technological Man. Prentice Hall. Karsenti, Eric. 2008. Self-organization in cell biology: A brief history. Nature Reviews Molecular Cell Biology 9 (3): 255–262. https://doi.org/10.1038/ nrm2357. Ladejinsky, Wolf. 1977. 52. Agriculture-some broader considerations. In Agrarian reform as unfinished business: The selected papers of wolf Ladejinsky, ed. Louis J. Walinsky, 462. Oxford University Press. Original edition, Economic Situation and Prospects of India—April 21, 1970. Lash, Scott. 2002. Critique of information. Thousand Oaks, CA: SAGE. ———. 2006. Life (Vitalism). Theory, Culture & Society 23 (2-3): 323–329. https://doi.org/10.1177/0263276406062697. Latour, Bruno. 2005. Reassembling the social: An introduction to actor-networktheory. Oxford: Oxford University Press. Levin, M. 2012. Morphogenetic fields in embryogenesis, regeneration, and cancer: Non-local control of complex patterning. Biosystems 109 (3): 243–261. https://doi.org/10.1016/j.biosystems.2012.04.005. Matthews, Eric. 1996. Twentieth-century French philosophy. New  York: Oxford University Press. Maynard-Smith, John. 1966. Sympatric speciation. The American Naturalist 110 (916): 637–650. https://doi.org/10.1086/282457. Moir, Cat. 2023. What is living and what is dead in political vitalism? In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T. Wolfe, 239–262. Cham: Springer. Petersen, Eric L. 2023. A ‘fourth wave’ of vitalism in the mid-20th century? In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T. Wolfe, 173–192. Cham: Springer. Russell, Edward Stuart. 1916. Form and function: A contribution to the history of animal morphology. John Murray: London. Shepard, P. 1967. Man in the landscape: A historic view of the esthetics of nature. USA: University of Georgia Press. Stjernfelt, Frederik. 2011. Simple animals and complex biology: Von Uexküll’s two-fold influence on Cassirer’s philosophy. Synthese 179 (1): 169–186. https://doi.org/10.1007/s11229-­009-­9634-­5. Sullivan, Heather, and Bernhard Malkmus. 2016. The challenge of ecology to the humanities: An introduction. New German Critique 43: 1–20. https://doi. org/10.1215/0094033X-­3511835.

28 

P. MCMAHON

Thompson, D. A. W. 1917. On growth and form. Cambridge: Cambridge University Press. Wagner, Gunter. 1999. Homologues, natural kinds, and the evolution of modularity. American Zoologist 36. https://doi.org/10.1093/icb/36.1.36. Webster, G., and B.C.  Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-­1750(82)91390-­2. Wessel, M. 1971. Fertilizer requirements of cacao (Theobroma cacao L.) in southwestern Nigeria. Amsterdam: Koninklijk Institut voor de Tropen. West-Eberhard, M.J. 2005. Developmental plasticity and the origin of species differences. Proceedings of the National Academy of Sciences of the USA 102, 1 (Suppl 1): 6543–6549. https://doi.org/10.1073/pnas.0501844102. Wolfe, Charles. 2011. From substantival to functional vitalism and beyond: Animas, organisms, and attitudes. Eidos 14: 212–235.

CHAPTER 2

The Debate: Cuvier and Geoffroy

Controversies in Early French Biology The good ship Dichotomy—Unity of Type vs Conditions of Existence— entered the Darwinian current by converting its terms from a debate about God’s primary mode of self-expression in nature to an argument about constraint and adaptation in evolution. (Gould 2002, 252)

Stephen Jay Gould was alluding to early controversies in the life sciences, particularly to the famous 1830 debate held in the French Academy of Sciences between Georges Cuvier (1769–1832) and Etienne Geoffroy Saint-Hilaire (1772–1844) on whether natural forms were determined by organism function, or were prior to function (Boucher 2015). In the nineteenth century a functionalism emerged that neither of the two protagonists of the debate subscribed to. Yet a synthesis of their ideas provided the foundation for a functionalist model, based on an adaptationism removed from the structural relations of the organism. Biology has turned to the gene, and its manipulation, as the fundamental, processing unit of life. A biology based on genetically directed adaptation to the environment provides the theoretical framework for gene manipulation. This would have seemed completely foreign to natural historians at the turn of the nineteenth century. Early ideas informing the new science of biology in the nineteenth century seem barely relevant now—scientists of the time appeared to be groping with completely different problems. Yet the very different positions of Cuvier and © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_2

29

30 

P. MCMAHON

Geoffroy were somehow integrated into a theory of evolution by Darwin (he referred to the work of both in developing his ideas on descent and adaptive change). The modern synthesis of Mendel’s (rediscovered) research and Darwin’s theory placed emphasis on the gene as directing function (rather than an integrated organism, which was Cuvier’s position). The neo-Darwinist school of thought maintains that biological organisms have emerged in history through adaptation to the environment via the mechanism of natural selection based on units of selection (genes and their expressed traits). Organisms are highly suited to their particular niches, a belief that harks back to Paleyism. William Paley, following a long tradition of natural theology, suggested God created living forms to suit their situation exactly, the best of all possible worlds (Ruse 1973). The debate between Cuvier and Geoffroy exemplified a shift in ideas of natural history from a picture of continuity in a rational order, to one of functional adaptation to conditions of existence. As biology became ‘modern’ in the nineteenth century, a new adaptive functionalism came to dominate life system models. This was not a straight-forward process, but involved an epistemic change from the rationalist thought of the eighteenth century (Foucault 1970). Charles Darwin’s interpretation of Cuvier’s discoveries demonstrated how heritable factors could enhance the adaptation of species within their particular conditions of existence. Following the integration of his theory with genetics and the new molecular discoveries of the twentieth century, these heritable factors were determined to be DNA, RNA and the gene. They became the new ‘reality’ of biology, the source of adaptation. The possibility of exchanging coded sequences between organisms and species became a real proposition, a fulfilment of the Baconian dream. Consequently, the biological justification for genetically transforming organisms has involved a process of reduction to their material and informational components. However, the stage for the emergence of biology in its modern form was set in the nineteenth century by Darwinian evolutionary theory and before that by debates within the French scientific community, including the 1830 showdown between Cuvier and Geoffroy.

Background to the Debate In 1795, the palaeontologist/anatomist Georges Cuvier joined the Museum of Natural History in Paris at the behest of Geoffroy Saint-­ Hilaire himself (Appel 1987). Both held chairs at the Museum, Cuvier’s

2  THE DEBATE: CUVIER AND GEOFFROY 

31

being specifically to oversee studies in zoological anatomy. Cuvier’s studies over the following 30  years underpinned a transition from a natural history that emphasised common forms to one focused on the adaptive function of anatomical structures. Nevertheless, Cuvier grew up in the eighteenth -century Age of Reason, an era of rationalist philosophy. The main aim of his anatomical studies of fossils and extant animals was to identify a rational structural design, a common plan underlying life itself; this was a project in keeping with the rationalist philosophy of his day. However, his careful anatomical studies uncovered different structural arrangements leading him to conclusions that conflicted with those of Geoffroy. These were debated in a series of public meetings at the French Academy of Sciences in 1830. Geoffroy’s own anatomical studies focused on the homology (or analogy in his terms) between different species. The strong focus on rational idealism in the late eighteenth century, especially in Germany, rejected notions of contingency: organisms were viewed to be united under one bauplan (or body plan). Toby Appel has given a detailed account of the disagreements, the politics and the personal conflicts within late Enlightenment French scientific circles, with a particular focus on zoology. By the end of the eighteenth century, she writes, the museum of natural history had become the premier institution for zoological studies, the envy of British and German scientists (many of whom visited, often through the generous accommodation of Cuvier). That such visits were made even in 1816 (by Robert Grant and others) the year following Waterloo, demonstrates the strong attraction of the museum to scientists in other countries and the openness of French biology at the time to all sorts of debate. Cuvier himself supported papers that were not entirely consistent with his own view, although he retreated into a more defensive conservatism in the 1820s (Appel 1987). The controversy underpinning the Cuvier/Geoffroy debate drew considerable attention bringing up questions related to metamorphosis and form. Johann Wolfgang von Goethe (1749–1832) the German poet, polymath and proponent of an archetypal morphology sided with Geoffroy in the debate (Appel 1987). To Michel Foucault (1926–1984) in his Order of Things (1970, 1966), the controversy highlighted a major transition: from Enlightenment natural history informed by rationalism and the Age of Reason, to a biology based on empirical observation, ideas on evolution and origin and an explanatory functionalism. It could be argued that both schools of thought permeate current biology: the homology that Geoffroy believed would reveal a unity of composition in all animals led to Darwin’s

32 

P. MCMAHON

theory of origin and descent and, more recently, to the discovery of regulatory genes common to vertebrates and invertebrates, while Cuvier’s organised structures morphed into a contingent adaptationism as the favoured explanation of evolutionary change. However, both could be considered as representing fundamentally different views of life to those held now, one of a rational order among living organisms or of structural relations within them (Webster and Goodwin 1982)—these aspects, central to early biology, have been lost in the new biology, a paradigm of genetic manipulation. Rational Order of Eighteenth-Century Natural History In the natural history of the Age of Reason, living and non-living components were ordered in a graded series of increasing complexity. An environment external to the organism was not the focus. Rather than an externality, there was continuum between the living and non-living. The world itself (including the ‘environment’) was within an order, one set by God. This rational order was active, progressive and had movement. Ideas on ‘transformisme’ were broached in France. But this was not a movement from an origin, it was not a historicity, since there was no origin other than the grand design of God, a blueprint or bauplan. The world operated according to this deistic blueprint. The line between the living and non-­ living was blurred because in the background hummed the mechanism of a smoothly operating clockwork universe, that included all entities. Therefore, the ‘environment’ worked in concert with the living. The concept of change in the rationalist biology of the eighteenth century was one of progression. In the Classical episteme, as Foucault called this period of thought, life forms were arranged according to visible structure (Foucault 1970). Linear series were proposed that were both synchronic and diachronic. The latter constituted pre-Darwinian ideas on evolution (or ‘le transformisme’ in French natural history). Foucault (1970, 172) suggested that discourse was closely linked to, even determined, how life was classified. The discourse of nature was a priori, based on order, preceding and predictive of natural history: “The table of signs will be the image of the things” (ibid., 73, emphasis original). Underlying this discourse was an assumption of continuity. Continuity in nature reduced classificatory schemes to specifics. The Comte de Buffon (1707–1788) proposed nature was a continuous series, rather than consisting of discrete groups, and any gaps found (apparent discontinuities

2  THE DEBATE: CUVIER AND GEOFFROY 

33

between forms) would later be filled with the discovery of the ‘missing’ intermediates (ibid.). His idea of species (or kinds) was influenced by Aristotle, proposing the only real ‘kind’ was a reproducing unit (the species) linked through time by generations: “Aristotle’s history of animals is today still perhaps the best we have in this genre”, he stated (quoted by Grene and Depew 2004, 74). However, Buffon also incorporated ideas of mechanism, wanting to explain “the hidden system of laws, elements and forces constituting primary, active and causative nature” (Greene 1981, 34). Underlying this philosophy was the idea of ‘matter in motion’. Unlike Aristotle, Buffon viewed natural history as non-teleological (hence life forms were not fixed and final), non-adaptive and uniformitarian (ibid.). Here Buffon clashed with Carl Linnaeus (1707–1778) who developed the system of classification we still use today. Linnaeus applied the categories of genus and family to groups (or species). To Buffon, genera and families were a product of the imagination, but Linnaeus regarded them as ‘real’. Linnaeus, unlike Diderot and others, rejected the idea of change. As a creationist, he believed the job of the natural historian was to describe and classify (ibid.). The genus was real and fixed, an eternal form. Yet the relation of discourse is apparent in Linnaeus’ taxonomy: his ‘real’ categories were broader than the characters themselves: Foucault suggested that the category, genus, ‘created’ characters (Foucault 1970). Progressive change was broached as an explanation for different life forms in the classical episteme. The great chain of being inherited from the Scholastics remained an influence. The Swiss naturalist, Charles Bonnet (1720–1793) proposed life forms progressed to greater complexity, the latter being closest to God. Change was invoked in this scheme. Jean Baptiste Pierre Antoine de Monet, Chevalier de Lamarck (1744–1829) also proposed progression from simple to complex life forms driven by vital material components. Although such evolutionary ideas were considered in the rationalist era, Foucault contends in his book that for Darwinian theory to be possible, a radical change was necessary: this changed natural history from a tableau of regular continuity to one of discontinuity between groups. Now, says Foucault, “the living being…tears itself free from the vast, tyrannical plan of continuities” (1970). The Linnaean scheme held fast in this modern turn, as a new category, the phylum (or embranchement), was introduced placing organisms with radically different structural arrangements into distinct and unrelated groups. Now any anatomical link between the octopus and the fish, for example, based on the rationalist idea of unity of composition, was rejected: the two were

34 

P. MCMAHON

recognised as being so different that there was no point in attempting to find common structures. Discontinuity in the natural world was real—the modern episteme had begun. Transition to Modern Biology In his ‘Order of Things’, Michel Foucault outlined the way in which biology in the modern episteme beginning at the end of the eighteenth century changed its focus from ordering the characters of organisms (notably the externally apparent characters of plants such as the number of petals and sepals, ovary position, which parts are fused etc.) to internal functions (mainly in animals) such as respiration, digestion and circulation: “The visible order, with its permanent grid of distinctions, is now only a superficial glitter above the abyss” (Foucault 1970, 273). Cuvier focused on zoology in his anatomical studies. Vertical relations were introduced between surface characteristics and deeper structures: “The internal link by which structures are dependent upon one another is no longer situated solely at the level of frequency; it becomes the very foundation of all correlation” (ibid.). The method of study changed from a deductive taxonomical approach to an inductive empirical one where the peculiar features of organisms were related to external environmental conditions. Rather than an array of characters providing the basis of a theory of prediction of similar characters, the focus changed to the individuality (or difference) of the anatomies of different species. Such anatomies reveal functions which in turn ‘reflect’ the conditions of existence. Kantian critique sanctioned “the withdrawal of knowledge and thought outside the space of representation” (ibid., 263). The relation between the beings and words of the classical systems was disrupted (ibid., 250). Discourse no longer provided a straight-forward representation of the natural world: “The very being of that which is represented is going to fall outside representation itself” (ibid., 260). As we entered into the modern episteme, the question now asked of a biological structure was what set of relations ensures that the organism is adapted to its conditions of existence. How is the carnivore structured for hunting, so that its internal organisation is appropriate for speed, for feeding and digestion and for perception. The great functions assumed a central place as comparative anatomy revealed the particular organic relations that adapted an organism to its needs. This began to undermine the strongly held belief (held in France and Germany) of a morphological

2  THE DEBATE: CUVIER AND GEOFFROY 

35

unitary type, a single plan, one that emphasised the common morphological features between organisms, rather than the adaptive roles of these structures. Focusing on this more traditional question of continuity and unity in nature, Geoffroy Saint-Hilaire had identified a unity of composition or homology across different species and taxa (Grene and Depew 2004, 151). An example of a homologous structure is the pentadactyl limb common to a wide range of taxa, from amphibians and reptiles to mammals and birds (Fig. 2.1). It denotes a common ancestor of tetrapods existing very early in the time of, or just before, the colonisation of land. In horses the middle (third) digit is enlarged to form the hoof, an instance of adaptation, yet the other digits are still present although in reduced form (Grene and Depew 2004, 151). Some species of pigs, for example, have four, rather than five, functional digits as result of secondary loss in

Fig. 2.1  The pentadactyl limb: top: salamander, toad, crocodile; bottom: bat, whale, mole, human. After Wilhelm Leche (1909) in Man, Origin, and Evolution. Available in the public domain. https://doi. org/10.5048/ BIO-­C.2013.3.f1

36 

P. MCMAHON

evolution. Geoffroy’s own work on embryology lent support to his ‘unity of composition’, due to the striking similarity in embryo structure between species. Les embranchements Cuvier’s research, rather than revealing a rational, continuity in living species and fossils, showed discontinuity between major groups of animals. He named these taxonomic groups ‘embranchements’, later to become our phyla. The four embranchements, Mollusca, Radiata, Articulata, and Vertebrata, were differentiated by their radically different structural arrangements or plans. Cuvier based this on the arrangement of their nervous system (Appel 1987, 3). However, these major taxa did have something in common. They all (despite their different structures) were admirably adapted to their particular conditions of existence. In other words, their structural arrangement (their anatomy) did not matter so much as their functional alignment with external conditions of existence. The great functions (digestion, circulation, sensation) are common to the fish and octopus and ‘work’ perfectly, but their respective anatomies are unrelated. To scientists just emerging from the Age of Reason (that Foucault called the Classical Episteme) these conclusions were challenging indeed. Rationalist biology was founded on an age-old scheme, the scala naturae or chain of being which arranged living forms from simple to complex. Cuvier (Fig. 2.2) upset this scheme with the discontinuity introduced with his four embranchements. Transformation over time was impossible Cuvier maintained. In each ‘type’, the structural arrangement was perfectly suited to cope with existence. His ‘Animal Kingdom’ surveyed these types, their anatomy and mode of living (Fig. 2.3). An alteration in even one part of this arrangement would result in the whole no longer being functional. Hence Cuvier was adamantly set against ideas of transformism, such as the theory put forward by Lamarck: life, Lamarck proposed, could transform itself within the framework of (divine) structural limitations. Cuvier’s fixism contrasted with more deistic accounts in rationalist biology (such as Lamarck’s). Geoffroy “a child of the Enlightenment” followed “a Deist view of nature”; God established laws “and nature was left to unfold in accordance with them” (Appel 1987, 7). Without the absolute dependence on an adapted structure that Cuvier endorsed, transformism would be possible according to this early evolutionary theory. According to John

2  THE DEBATE: CUVIER AND GEOFFROY 

37

Fig. 2.2  Georges Cuvier holding a fossil of a fish (Wikipedia Commons)

Greene, Cuvier’s orthodox beliefs returned natural history to an Aristotelian fixism: Cuvier’s … comparative anatomy and Aristotelian functionalism, far from overthrowing the static paradigm of natural history, served only to strengthen and further articulate the taxonomic, teleological approach to natural history. (Greene 1981, 38)

In the turn towards an empirical biology, the relation of the inner functions of organisms, especially animals, to their environment was derived increasingly by an inductive method. Functions and associated structures were linked to conditions of existence. An empirical description of a bird’s

38 

P. MCMAHON

Fig. 2.3  A page from Animal Kingdom by Georges Cuvier

beak, for example, was used to provide an account of its feeding behaviour. Anatomical studies indicated that the digestive tract of mammalian carnivores, such as dogs and their relatives, was short compared to that of herbivores, such as cows or sheep, and this was directly linked to their different modes of nutrition. Herbivores digest cellulose (with the help of symbionts) and need longer digestive tracts than carnivores. Therefore, anatomical studies conducted by Georges Cuvier and his contemporaries revealed the way that internal anatomical arrangements of organisms reflect their conditions of existence (Foucault 1970). In this way, Cuvier was a key transitional figure in the epistemic shift proposed by Foucault. While he was concerned with identifying common plans, applicable to all animals, the holy grail sought at the Museum, his close anatomical examinations revealed inconsistencies that could not uphold a unitary type or single structural plan. Furthermore, Cuvier was a Protestant in a mainly Catholic country, and was not as wedded to scholastic notions of the chain of being or a unitary plan as some of his colleagues, holding “the traditional belief that the Creator could not be constrained in His activity”

2  THE DEBATE: CUVIER AND GEOFFROY 

39

(Appel 1987, 7). The four plans Cuvier identified were created, divine plans; they were separate entities with no transitional forms. But each embranchement had countless variations within, accounting for the diversity and richness of the living world. For Cuvier, the guiding principle in the anatomical arrangements of animals was their ‘Conditions of existence’, which was supported by the ‘Correlation of parts’ and ‘Subordination of characters’. This can be contrasted with Geoffroy’s guiding principle, Unity of Composition, based on the ‘Principle of connection’ and ‘Balance of parts’ (Grene and Depew 2004, 141). Cuvier’s Conditions of existence might be better termed ‘Prerequisites of existence’, since he was primarily concerned with the internal structural arrangements of his animal specimens, viewing these arrangements as perfectly suiting the living animal to its niche (ibid.). Rather than environmental conditions, “what concerned him first and foremost was the integrated, harmonious coordination of all the parts, each functioning to produce a functioning whole” (ibid., 139). Thus, Cuvier introduced a systemic element. He was critical of the grandiose system-building of Lamarck, believing that conclusions should be derived from collected data, yet his organism was a hierarchical system ruled by the need for survival, followed by the major organs and then minor (secondary) characters and structured by the relations between parts. Hence, different structural forms were united by the great functions and the needs of the individual. Cuvier transposed a form based on common structures (held to in the era of rationalism) to one based on function. In a new hierarchy, function became the ruling order. The suitability of the organism to its environment was an Aristotelian view that influenced Cuvier. But he was also influenced by Linnaeus and his classification scheme, adding to the Linnaean scheme by introducing a new category, the phylum with no structural relation to other phyla (Appel 1987). Cuvier’s adapted organism as the final cause was teleological, and the product of divine intervention. Cuvier introduced a new hierarchy to the relations that constitute the organism. In his scheme. secondary structures at the periphery were dependent on less visible primary structures (such as the quadruped arrangement—four limbs—of vertebrates), and, at an even deeper level, the ruling functions. Only a few functions were responsible for a “visible diversity of structures” (Foucault 1970). These functions were less obvious, even invisible, a hidden “functional homogeneity” (ibid.). Relations were now foundational, rather than something to be recorded. Now parts

40 

P. MCMAHON

(organs) were interdependent, forming “a single system” (ibid.). The hierarchy was characterised by the unchanging, constant features of the organism (like the quadruped structure) leading to the more flexible (wing and limb structure) secondary structures. Functions were also placed in a hierarchy: for example, circulation, a feature of higher life forms versus digestion, found in all animals. The deeper one delved, the more abstract relations became, yet more primary to the hierarchy. Multiple structures were derived from “the great mysterious, invisible focal unity” (Foucault 1970). Foucault proposed that new transcendental objects (like Cuvier’s organism) were not restricted to biology, but also featured in nineteenth-­ century linguistics and economics, launching a new (modern) approach to epistemology. Comparative methods of study came to the fore at this time. The abstract relations determinant of these transcendental objects seemed to be unfathomable, akin to Immanuel Kant’s ‘things-in-themselves’ so that science began to focus on positive, recordable data as “it is in the division between the unknowable depths and the rationality of the knowable that the positivisms will find their justification … and [thus] positivism in the sciences is linked to a transcendental philosophy (first put forward by Kant but extended to transcendentals situated with the object)” (Foucault 1970, 265). Geoffroy Put Morphological Structure First The approach to evolutionary studies pioneered by Geoffroy (Fig.  2.4) prioritises morphology; in fact, form, or structure is seen to restrict the possibilities of adaptive evolution. For example, the flippers of dolphins and whales (cetaceans), are adaptations evolved from a terrestrial ancestor for an aquatic existence, while bat wings have been modified for flight (Fig. 2.1). The pentadactyl bone structure of the flipper or the bat wing is essentially the same as that of ancient amphibians and, as the evidence provided by homology suggests, also of a common plan in all tetrapods. If the whale flipper and the bat wing are homologous structures, found in all tetrapods and modified for a diversity of uses, they cannot be ‘perfect’ adaptations. The structure is prior to its function. Homology is also identified in the four limbs in tetrapods evolved from appendages found in all jawed vertebrates; jawed vertebrates including fish have two pairs of fins (dorsal and pectoral) or two pairs of limbs, whether functioning or vestigial, and no more. Structural constraints also are definitive in the class of

2  THE DEBATE: CUVIER AND GEOFFROY 

41

Fig. 2.4  Etienne Geoffroy St Hilaire (Wikipedia Commons)

insects; insect species have no more than six legs—would more legs have adaptive use as found in the arachnid class (spiders, mites and other groups) which have eight? It is likely that insect ancestors had more than six legs but some of these were transformed into other structures. This fits ideas of transformism of the time: existing elements were used for different purposes. Some essential differences emerge in the framework by which the rationalists viewed the organism. The model still predominant in the early nineteenth century was superseded by an environmentally defined model, developed by Cuvier in his paleontological and anatomical studies. To Cuvier the relations between the great functions within organisms and their external conditions were invariant. Influenced by Aristotle’s ideas on the species as a functional unit (Greene 1981), this in effect was ‘form’ for Cuvier, the invariance that defined the organism. The structural arrangement and the structures themselves were secondary. In this way they were contingent to the rule of function: different arrangements are possible, as long as they serve their function. His colleague, Geoffroy, who also applied

42 

P. MCMAHON

methods of comparative anatomy, disagreed saying different life forms were linked by common morphologies. Geoffroy’s insistence that all living forms were united under a single plan was perhaps prescient to some modern discoveries, particularly in evolutionary developmental biology (evo-­ devo) (Le Guyader 1998 [2004]). The relative order of elements, such as bone structures, was considered to be invariant (the principle of connection) but could be transformed for different purposes, while function (the relation to external conditions) was secondary. An example of this principle is Karl Reichert’s discovery that the ossicles of the mammalian middle ear can be identified in reptiles as part of the jaw (the articular and quadrate bones) (Webster and Goodwin 1982). Differing from Lamarck’s proposal of gradual change, Geoffroy proposed that transformational changes occurred in bursts (Appel 1987), an early suggestion of saltation (sudden changes leading to new forms) as an evolutionary mechanism. His interest in teratology contributed to this view (Appel 1987, 134). As opposed to Cuvier, he believed the environment directly affected ontogeny (including the development of monstrosities), speculating that this could contribute to evolution. A different view of form is presented: rather than being Aristotelian it has more in common with mechanism, the philosophy of the Age of Reason that challenged Aristotelian explanations. Form as Homology The different positions defended by Cuvier and Geoffroy, rather than being clear-cut scientific hypotheses, carried with them the presumptions of earlier philosophies. The particular was opposed to the universal in their debate. The ‘form’ defended by Cuvier is very individual, based on functional relations to particular conditions of existence. In contrast, Geoffroy’s form was a universal and transcendental proposition. Mechanical ideas were adopted by Geoffroy: a law of balance and compensation was based on the limited availability of materials in living structures: “Nature works constantly with the same materials. She is ingenious to vary only the forms.” He adopted Johann Goethe’s and Carl Friedrich Kielmeyer’s law of compensation that proposed the parts of an organism that needed more nutrients, blood supply or resources were compensated by reduced supply to other parts (Rieppel 1990). All adjustments were internal, without environment input. If one structure was enlarged in the arrangement, others were reduced. A horse’s hoof is one example of this, as it is formed

2  THE DEBATE: CUVIER AND GEOFFROY 

43

from one enlarged toe, while other toes have been reduced. This contrasted with Cuvier’s idea of adaptation according to need. If a structure was needed it was produced. Materials to supply this need were not limited. He scoffed at the idea that some parts are reduced in compensation for others. The adaptationist view, that prevails now, in which organisms have evolved specialised features to cope with their conditions of existence tends to overshadow the view that homologous structures are not totally plastic, although they could be modified. To Richard Owen, who believed morphological structures, such as the pentadactyl limb, are permanent and everlasting, modifications of such structures for particular purposes (flying, swimming etc.) created an ‘adaptive mask’. After the debate and Cuvier’s death shortly after, biologists turned their attention to the striking homologies made evident by studies in anatomy, as well as new ideas of history and origins. Despite Cuvier’s superior presentation of data and predominance in the debate, it was Geoffroy’s morphological approach, further developed by his son Isidore, that was taken up in Britain by Richard Owen and others. Cuvier’s conviction that common morphological structures between organisms must lead to them being ill-adapted, so that they could only impair function, became less influential (Appel 1987, 5). Although, transformism remained unpalatable to most biologists, Geoffroy’s ideas had established the framework for a possible theory of morphological transformations within a unified type. This set the stage for Darwin’s theory of descent (although as Foucault pointed out, Cuvier’s ideas were adopted in his account of adaptation). After Darwin, the pendulum swung to the other extreme so that natural history appeared to be focused exclusively on the diversity and differences between living forms rather than their common properties or homologies. Working together in their earlier careers, both Cuvier and Geoffroy were interested in developing workable systems of taxonomy based on natural kinds, and not on preconceived notions such as the Great Chain of Being. However, Cuvier was already leaning towards differentiating types based on their primary functions, and how they fitted into their ‘conditions of existence’. Later, Cuvier entered a period of ascendency in French scientific circles with the support of Napoleon, becoming a man of many hats. Geoffroy, although a member of Napoleon’s Egyptian expedition (or maybe because of his absence from Paris) lost influence in the French capital. As Toby Appel outlines, Cuvier was extremely political exerting his influence through recommendations and even implied threats regarding

44 

P. MCMAHON

further employment. He actively promoted his embranchements as absolutely and irreconcilably separate and gave precedence to the functional anatomy that described them (Appel 1987). Nevertheless, Geoffroy never accepted that there were no linking life forms between Cuvier’s embranchements. Further, his studies revealed that common structures could have different functions; these studies led to the differences between the two that came to a head in the 1830 debate.

The Debate at the Academie des Sciences In the 1820s, the rift between the views of Cuvier and Geoffroy widened into an apparently irreconcilable gulf. In 1830, Geoffroy who had pressed for years for the issues to be debated in a scientific forum (Appel 1987, 144) was rewarded with the arrangement of a formal debate by the French Academy of Sciences. The debate was undertaken in a number of sessions. The point of contention was “whether animal structure ought to be explained primarily by reference to function or by morphological laws” (ibid., 2). Seating overflowed in these sessions and the debate was reported in the press and discussed widely in Britain and Germany—in the latter country Goethe was to write two articles on the debate. It became clear that the gulf separating the two scientists was not only based on available evidence, but as Geoffroy stated in a memoir: “It is a question of philosophy that divides us” (Appel 1987, 152). The disagreement extended to scientific style, personalities and social context, and the subtext recognised widely was the different views on evolutionary change held by the two opponents. Therefore, the debate was later interpreted “as a pre-­Darwinian testing of the doctrine of transmutation of species” (ibid., 3). The main point of disagreement was that, according to Geoffroy, taxa showed homology across widely different organisational structures, such as molluscs and vertebrates. Of particular concern to Cuvier was the claim by Geoffroy and his followers that homology could be demonstrated between mammals and insects. He presented evidence of how vertebrate and invertebrate structure was fundamentally different and unrelated. Geoffroy countered with his idea that in insects the central nervous system of the vertebrates (located dorsally) was flipped to a ventral position. By imagining major transformational changes, a homology between the two groups could be shown. The general consensus was that Cuvier’s arguments prevailed in the debate. He used a wealth of empirical data to undermine Geoffroy’s claims, while the latter was more concerned with general principles, expecting supporting data to emerge in future research. In some ways the

2  THE DEBATE: CUVIER AND GEOFFROY 

45

ideas of Geoffroy have been vindicated with the demonstration of a deep homology across phyla (including insects and mammals). Homologous genes found in vertebrates and invertebrates such as the Homeobox (Hox), group are concerned with regulation of developmental genes. These transcription factors bind to DNA determining major pathways of development including dorsal-ventral orientation and cell shape. In the early twentieth century, during the ‘eclipse of Darwinism’, Edward Russell recounted the debate in his Form and Function (published in 1916). Recognised as a Lamarckian, Russell favoured Cuvier in the debate despite the latter’s stance against transformism. This was because, according to Toby Appel, “Cuvier had given precedence to holistic functional considerations rather than strictly morphological considerations” (p. 3). Russell saw, in common with Lamarck, a teleology (a purpose) in Cuvier’s ideas, absent from Geoffroy’s morphological series. As discussed later (Chap. 6) this was also the time Saussure developed his structuralist linguistics, which was a critique of the ability of diachronic (evolutionary) linguistic models to account for function. If organicism is based on a whole as supported by each part, structuralism places the emphasis on the relations between parts to constitute a functioning whole. Cuvier’s system has common features with the structural linguistics of the twentieth century. Like Saussure’s langue, his system was static. It did not countenance change, particularly incremental change. Darwin was to bring that in later with natural selection of variable traits. To Cuvier, the whole organism, its structural arrangement, the relations between its parts enabled its survival. His anatomical studies showed that “in any organism, the parts cohere in a meaningful, well-adapted whole” (Grene and Depew 2004, 135). Rather than parts working in concert to support the whole, the whole was its set of relations. This comes much closer to Saussure’s system, a functional entity, based on the relations between its constituent elements. There are no independent parts ‘contributing’ to the whole in Saussure’s system. Equally, there are no autonomous parts in Cuvier’s organism: the ability of the organism to function under its external conditions is based on its internal ‘correlation of parts’. Le transformisme The theme underlying the debate was Lamarck’s transformism (although this was not the explicit topic of discussion), which Geoffroy supported and Cuvier opposed. Lamarck (like Cuvier) was also interested in the relation of organisms to their environment, but he took a different point of view: organisms adapted to their needs by struggle and effort, and the resulting changes

46 

P. MCMAHON

were passed on to progeny. Cuvier rejected the notion of incremental change over time as a response to the environment. This might seem contradictory since the environment (conditions of existence) was central to Cuvier’s adaptation thesis. But to Cuvier, species were already adapted to their environments. In his view, conditions of existence were central to any explanation of biological form, but that form was already in place. To Lamarck the environment was not a driver of evolution, but rather evolution was a vital process internal to life forms as organisms strove for greater complexity. Yet biological diversity needed an explanation and this is where Lamarck invoked his ideas of adaptation according to need. Depending on external conditions and the needs of the organism, organisms could transform themselves into a better adapted phenotype. His was a thesis of self-transformation. The internal striving of the organism to adapt to its conditions explained diversity (the branching of taxonomic forms at each level in his progressive scheme of simple to complex types). Paradoxically, he did not believe in extinction, while Cuvier, the fixist, demonstrated that the fossil record showed types that no longer existed (Grene and Depew 2004, 151). Cuvier’s explanation of adaptation was one akin to creationism—the perfectly adapted species was a given. His notion of change (to account for species extinction, which his research had proved to be a real event) was ‘replacement’. Following a major catastrophe leading to extinction, species were replaced by new forms. Each replaced species was proposed to be adapted to its new environment. In contrast, Geoffroy’s notion of transformism has similarities to Lamarck’s, one of an internal adjustment—but rather than gradual change, he envisaged more radical transformations. More recently, evolutionary developmental (evo-devo) biologists have addressed phenotypic change as a possible evolutionary mechanism, without necessarily invoking a Lamarckian explanation (see Chap. 7). Ideas such as modular gene expression systems that can be triggered by environmental variables are ­raised. Gerry Webster and Brian Goodwin (1982) contrast the rationalism that informed 18th to early 19th century French natural history, with Darwinian theory. The former was based on a concept of rational continuity, while the latter was based on a contingent adaptation. Now formalist ideas, whether Geoffroy’s or Cuvier’s, seem obsolete in modern biology. Their account endorsed an immanent form, a feature of the organism that was dropped in later theory. The common ground between the two protagonists is emphasised by Webster and Goodwin—both endorsed a structural ‘whole’, although Cuvier’s model was more akin to a functional system. Similarities can be discerned in Saussure’s langue (see Chap. 6). There was

2  THE DEBATE: CUVIER AND GEOFFROY 

47

no room for diachronicity in Cuvier’s organism. Cuvier’s catastrophe theory served to explain extinction and the fossil record. Yet clearly his ideas had no room for more continuous changes over time, and in fact rejected evolutionary change. However, early ideas in biology continue to live on in biological structuralism, although these positions are marginalised, even ridiculed. It is important to delineate what these ideas are and how they differ from, even challenge, the current neo-Darwinian paradigm. Key points they raise are a rejection of both atomistic and idealist concepts of holism, which they propose amount, in the end, to the same thing: “At least as far as these biological phenomena are concerned, atomism and holism are mutually constitutive concepts; they are two sides of the same coin” (Webster and Goodwin 1982, 21). Both lead to a mechanical conception of the organism. In idealism the ‘idea’ is a non-material conditioning of the organism (which is effectively a machine) while the atomistic explanation is that of an assembly of parts that is determined by a central genetic program associated with the germplasm – the parts again work as a machine. The structuralist alternative is that life is based on a plan or rational morphology. While such a rational morphology could be transcendent or immanent, the structuralist position is one of immanence. Webster and Goodwin follow Kant’s conception of the organism as a self-organising entity. An immanent set of relations could be considered a ‘set’ of possible transformations. Webster and Goodwin themselves conflated very different views expressed in early French zoology into one seamless proposal of the organism functioning not by the action of its different parts but how these operate together in concert. But these alternative models have drawbacks, in particular inadequate explanations of change (or no change, as proposed by Cuvier, other than that introduced by creation). This deficiency was addressed by Darwin, who built upon Cuvier’s functionalism, rejecting his model of perfect functional integration. In its place, he introduced the notion of incremental singular changes in response to environmental changes that improved function (or adaptation). He also rendered function as relative, with room for improvement (Grene and Depew 2004, 212). The transcendental archetype was another problem, an object that escaped any empirical verification, and this was also addressed by Darwin, turning it into an historical product. Yet Webster and Goodwin ask: Is the putative ‘ancestor’ really any different to the archetype? They propose that general structuralist principles can be drawn on to explain biological phenomena.

48 

P. MCMAHON

The comeback of formalism at the end of the 19th century, during the period described as ‘the eclipse of Darwinism’, can be traced to these earlier French rationalist ideas. However, divisions occurred on the principles underlying form (Cuvier’s organism or Geoffroy’s unitary composition). Also, at times structuralism has been conflated with organicism – but Archie Bahm, for example, defined them as separate systems (Bahm 1983). Structuralism is concerned with systems of relations that define the properties of its elements, while organicism is characterised by a polarity and reciprocal tension between the parts and the whole of a system. In organicism the parts (e.g. the organs, the genes) are functional in their own right, but also contribute to the production of other parts. Other systems described by Archie Bahm, include atomistic, cybernetic and emergent systems. All of these have been invoked in biology.

References Appel, Toby. 1987. The Cuvier-Geoffroy debate. Oxford University Press. Bahm, Archie J. 1983. Five systems concepts of society. Behavioral Science 28 (3): 204–218. https://doi.org/10.1002/bs.3830280304. Boucher, Sandy C. 2015. Functionalism and structuralism as philosophical stances: van Fraassen meets the philosophy of biology. Biology & Philosophy 30 (3): 383–403. https://doi.org/10.1007/s10539-­014-­9453-­z. Foucault, Michel. 1970. The order of things: An archaeology of the human sciences. New York: Pantheon Books. Gould, S.J. 2002. The structure of evolutionary theory. Cambridge, MA: Belknap Press of Harvard University Press. Greene, John C. 1981. Science, ideology and world view. University of California Press. Grene, Marjorie, and David J. Depew. 2004. The philosophy of biology: An episodic history. New York: Cambridge University Press. Le Guyader, Hervé. 1998 [2004]. Etienne Geoffroy Saint-Hilaire, 1772–1844: Un Naturaliste Visionnaire, (Un Savant, Une Époque). Paris: Belin. Translated as Étienne Geoffroy Saint-Hilaire, 1772–1844: A Visionary Naturalist, 2004. In Chicago: University of Chicago Press. Rieppel, Olivier. 1990. Structuralism, functionalism, and the four Aristotelian causes. Journal of the History of Biology 23 (2): 291–320. Ruse, M. 1973. The philosophy of biology. London: Hutchinson and Co (Ltd). Webster, G., and B. C. Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-1750(82)91390-2.

CHAPTER 3

Setting the Stage for Evolutionary Theory

Paradoxes Arise in Pre-Darwinian Theory The Cuvier-Geoffroy debate highlighted some paradoxes. Geoffroy’s interpretations of homology supported an evolution (‘le transformisme’) based on a single form (plan of life), while Cuvier isolated forms so different that it was impossible to countenance any structural continuity between them, discounting the possibility of incremental change. Yet, while Lamarck’s evolutionary theory and Geoffroy’s ‘Unity of Composition’ principle inspired Charles Darwin’s inquiry, it was the arch-­ conservative religious beliefs of Cuvier, out of tune with the materialist ideas of the Enlightenment, that formed the basis for Darwin’s own theory of natural selection. Despite Cuvier’s resistance to ideas of transformism, Michel Foucault maintains that it was Cuvier’s ideas (rather than Geoffroy’s or Lamarck’s) that paved the way for Darwin’s theory of evolution by adaptive change through natural selection (Foucault 1970). Darwin turned to Cuvier’s discoveries to develop his own theory, but it is important to recognise that Cuvier himself insisted on the invariance of the organism and fixicity of species. He based his views on Aristotle’s biology: form was intricately linked to function. On the other hand, the rationalists led by Lamarck and Geoffroy endorsed ideas of morphological change within the rational, deistic order. These scientists were the progressives of the age, discussing ideas of evolutionary change, dangerous ideas on epigenesis and encountering strong resistance from orthodox © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_3

49

50 

P. MCMAHON

religion. Therefore, Foucault suggests it could only have been someone like Lamarck who could develop an early evolutionary theory, but impossible for someone holding Cuvier’s beliefs. Yet, paradoxically, it was Cuvier’s science that held sway in the theory later developed by Darwin. Cuvier, the fixist, inspired Darwin’s proposal of evolution by natural selection, while Geoffroy, the deist, was more open to possibilities of change or evolution as part of the rational order put in place by God. This idea of a clockwork universe had been around since Newton. Hence, Darwin was initially drawn to Geoffroy’s account of homology in developing his theory of evolution. While the rationalists viewed the world (organisms and environment) as ordered in a rationally derived continuum (to the extent that if gaps appeared in the continuity of life they were left blank to be filled in by future discoveries), Cuvier’s studies revealed that such a continuum was spurious and developed a thesis of intransigent conditions of existence against which God had created perfectly functioning organisms. The great functions united the very dissimilar forms identified by Cuvier, such as those of vertebrates and molluscs. The life/non-life boundary was deepened into a moat, in contrast to the tables of Enlightenment taxonomists in which the boundary between life and non-­ life forms was blurred: “Nineteenth century nature is discontinuous exactly in so far as it is alive” (Foucault 1970). Cuvier’s empirical investigations revealed organism function that matched external conditions of existence—the inner organism mapped or mirrored its external conditions. Hence life and non-life (or potential death) did hold a relation but as a reflection—the organism reflected or represented its environment. However, the search for God’s rational plan for life in the natural history of the eighteenth and early nineteenth centuries did not focus on function but the bauplan, an elusive unitary anatomical plan providing the blueprint for animal forms. Revealing this plan was the most important task for these scientists, including Cuvier in his early studies. Cuvier’s genius was to uncover not a single plan, but four embranchements, all perfectly functional in their respective environments, but also completely different: “It seems not unfair to say that we see Cuvier working to find unity within separation, and Geoffroy to conquer separation by discovering unity” (Grene and Depew 2004, 139). Darwin, subsequently, took up the ideas of “the illustrious M. Cuvier” (Darwin 1859) rather than those of Geoffroy in the formulation of his theory of evolution by natural selection. Further, Cuvier was the first to prove extinction was a real phenomenon. He suggested regular catastrophes, particularly great floods, could

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

51

explain the disappearance of species. Their ‘replacement’ with changed structures could only be accounted for through intervention, presumably divine. This explained the fossil record to Cuvier. Earlier creations, he proposed, had been wiped out by catastrophes, such as great floods, but an interventionist event replaced these with a new set of life forms. In contrast, Geoffroy (following Lamarck) supported early ideas of evolution, the notion of ‘le transformisme’. A creationist explanation was removed in the Darwinian model. Change was postulated as a result of variation among progeny of species and the survival to reproductive status of the best adapted. Thus as the environment presented new challenges a response was possible from life forms due to the natural variation among progeny (now known to be due to genetic differences), including novel promising characteristics (mutations) that were heritable. Many more progeny were born than survived (Darwin turning to Thomas Malthus for this account). Natural selection occurred by weeding out the poorly adapted—the best adapted survived passing on their characters to the next generation. Therefore, evolution proceeded through a cycle of life and death, not unlike Cuvier’s catastrophic theory of change, but repeated in each generation, thus accounting for gradual change. This explanation sharply contrasts with proposals of evolution in the rational order by Lamarck. Lamarck did not accept extinction as a real event. The central drive, Lamarck proposed, was towards an increase in complexity—accounting for the advance from simple to higher organisms. Branches from this linear progression were explained by Lamarck as an internal modification (a tweaking of the design God provided) as environments changed; in his use/disuse theory, use strengthened a character that could also be inherited, while disuse weakened the relevant character (as in the wings of penguins, which lost their ability for flight). These modified characters were passed on to progeny. The modern episteme was informed by a curious mixture of traditional orthodoxy and new methodologies, such as comparative anatomy. The conservative Cuvier, an inspiration for Darwin’s theory, remained firmly within the theological framework of the time, eschewing materialist accounts, and insisting on the fixicity of species, while more progressive scientists toyed with evolutionary ideas, yet remained attached to explanations of a tabular continuity put forward in the rationalist era (or classical episteme). In addition, Cuvier remained an Aristotelian, even in the era of Mechanism, which was the key philosophy underlying the Enlightenment.

52 

P. MCMAHON

Therefore, unlike Lamarck and Geoffroy, to Cuvier species were teleological (a sort of ‘final cause’) and fixed, not beings in the process of structural change.

Demise of Formalist Biology As Michel Foucault (1970) proposed, a new Darwinian synthesis was made possible when empirical investigations by anatomists such as Georges Cuvier were combined with ideas of evolutionary change, promoted in geology by Charles Lyell (1797–1875) and in biology by Lamarck among others. The influence of Georges Canguilhelm can be discerned in the thought of his former student: for new scientific thought to develop, the right conditions for its reception must be in place. Neither the relation of anatomy to survival of individual organisms under particular environmental conditions nor the proposition that species change could alone provide an adequate evolutionary theory. Rather, both were required before thinkers such as Darwin and Alfred Wallace could produce an explanation of evolution formidable enough to challenge the contemporary religious dogma. As Canguilhelm and Foucault pointed out, science does not follow a path free of its social and historical setting. Richard Owen, in his earlier studies, could not accept the notion that species change, only to change his position later as evolutionary ideas took hold. The life sciences of the eighteenth century were transformed in the nineteenth century because the conditions of possibility of knowledge differed between these periods. The effect of conditions of knowledge on the scientific belief held to in a particular historical moment is even more obvious in the case of scholasticism. Such concepts as the chain of being and immutability of matter and species were unquestioned and unquestionable. The dominant discourses of the time, therefore, did not allow ideas such as mutability or evolution to develop. The disagreement between Cuvier and Geoffroy Saint-Hilaire, in particular, concerned the importance of adaptations, which to Cuvier was a reflection or an outcome of an organism’s conditions of existence; in other words, structures appeared to be mainly adapted to the particular needs of organisms in relation to the environment they inhabited. Cuvier therefore diverged considerably from his former rationalist (background, beliefs) and endorsed a view later developed into an historical account of adaptationism. Even though Cuvier, as a fixist, did not accept evolutionary change, an idea that had been broached for a number of years by Lamarck

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

53

and others, he represented a transition in thought from a morphological to functionalist position (Foucault 1970), while retaining beliefs in the immutability of species and a grand design or plan from which they emerged. The adaptationist view (which Cuvier unwittingly set in motion) is that organisms are malleable; forged to suit external environmental conditions, their anatomies directly reflect their particular mode of existence. Cuvier collated evidence that supported the eminent suitability of the organic structures in animals for their particular requirements for survival. This suitability was presumably the result of divine intervention, although he did not explicitly suggest this. But the explanation fitting his Protestant beliefs was that species were created by God as exact fits for their environments. Nevertheless, despite its emphasis on adaptation, Darwin’s theory moved markedly away from Cuvier’s ‘complete’ structures: [U]nder the guidance of natural selection, adaptation seems to return in a thoroughly non-Lamarckian fashion. It seems even more emphatically anti-­ Cuverian, since for Cuvier it is the whole integrated organism that is, now and forever, adapted to its proper niche in nature. (Grene and Depew 2004, 151)

While Foucault contends that Darwin took up Cuvier’s notion of adaptation and built his theory of natural selection around it, Marjorie Grene and David Depew point out that Cuvier’s functional organism was a ‘whole’ and, to Cuvier, changing parts incrementally would interfere with this functional structure. Darwin’s interpretation of adaptation was different in that he proposed that distinct units of variation, selected through environmental pressures, could enhance adaptation. Therefore, the formalism of early biology needed to be overturned in order to make way for the new evolutionary theory. Evolutionary theory today is still predominantly Darwinian, proposing that evolutionary change occurs through particulate responses to environmental conditions.

An Atomistic Paradigm The Darwinian account of homology is that life forms have evolved from a common ancestor. This sits well with the empiricism favoured in modern biology, for such an ancestor must be a real (if rather vaguely delineated)

54 

P. MCMAHON

life form, and not a rationalistic plan or structure. But the main theme in the biological revolution of the nineteenth century was the adaptation of organisms to a range of environmental conditions, leading to the huge diversity apparent in living natural forms: The problem of form is thus effectively ‘reduced’ to the problem of functional adaptation and its specificity. With the eventual acceptance of ‘Darwinism’, this conception of the problem became that of mainstream biology and all other aspects tended to be ignored. (Webster and Goodwin 1982, 24)

Darwin, embedded in the British tradition of empiricism and utilitarianism, as well as Paley’s protestant theology and ideas on limits propagated by Malthus, was concerned with the “(concrete) trees” rather than the “(abstract) wood” (ibid., 22). In contrast, rationalist biologists of the Enlightenment, including Cuvier, aimed to “discover the real unity or ‘form’ hidden in the diversity of appearances which constitute the ‘content’” (ibid.). With the emphasis on an empirical approach that revealed difference and diversity, the structural and physical limitations that organisms are subject to and the importance of structural relations between the ‘parts’ that make up an organism tended to be bypassed or discounted. Therefore, the wood could no longer be seen for the trees since any attempt at a unifying theory of structure, such as that pioneered by Geoffroy, Cuvier and others was frowned upon and biology became an unending checklist of facts and observations: This sort of approach naturally tends to emphasize the diversity of the biological domain, the specific peculiarities of organisms and the way that species differ from each other in relation to their particular environments and modes of life. It thus encourages an ‘atomistic’ rather than a ‘systematic’ conception of the totality of organismic forms; a conception which is at one with the conception of the individual organism as an ‘atomistic’ or ‘mechanical’ aggregate of parts each with a primary functional relation to the external world. Thus at both the individual level and at the level of the ‘totality’ there is a tendency to see the biological domain as irreducibly complex and ‘given’. (Webster and Goodwin 1982, 22)

Cuvier’s thought is aligned with the rationalist idea of unity of form (and indeed his career was concerned with demonstrating a unitary concept of the organism). Foucault’s analysis portrays a picture of transition

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

55

embodied by Cuvier’s work, between the search for a rationalist morphology and an emerging emphasis on functionalism. Nevertheless, Cuvier’s form was developed within the Kantian concept of the organism. He reminded his readers of “Kant’s insistence that we refer to the whole organism” (Grene and Depew 2004, 139). Immanuel Kant contrasted life to machines: machines demonstrate functional unity since each part only exists for the other parts in the system, so that each part is a condition of the function of the other parts, leading to a common end. An organism (life) is structured differently: each of its parts exists for other parts (like the machine) but also by means of each other, each part being involved in the production of the other parts. In other words, Kant proposed the organism had both functional and structural unity. This contrasts markedly with other ‘idealist’ concepts in German biology. Unlike Kant’s idea of “the whole as whole” (Webster and Goodwin 1982, 21, italics original), a self-organising actuality, Schelling and others proposed an ‘Idea’ directed life, which was otherwise mechanical, a functional unity. The active agent was a ‘spiritual organising centre’: The organism is, therefore, the phenomenon of its ‘Idea’, it is no longer conceptualised as a ‘self-organising totality’ but as an ‘expressive totality’, an expression of the nature and activity of the ‘Idea’. (Ibid., 21, italics original).

Therefore, German biology was characterised by a dualism: between the spiritual and material, and between a directive organising centre and mechanical organism. Transformism (the rearrangement of existing parts in an evolutionary process) becomes difficult to support in the German archetypal model. However, to Webster and Goodwin (1982) the archetype fits well with the Weismann barrier and Central Dogma of Francis Crick: in other words, the spiritual organising centre of German biology reappears as the gene and germline in modern biology. This could also be considered as a separation of form and matter: The way in which this conception was inscribed in twentieth century biology and which refers to the hierarchical distinction between reproduction and metabolism—that is to say a genome conferring on living beings their distinction, i.e., their form, and a metabolism ensuring the continuous material production of this form—contributed to reinforcing a vision of living beings that is both ‘genomecentric’ and metabolically egocentric. From this point of view, the classical question of the priority of the genetic or the

56 

P. MCMAHON

metabolic, of form or matter, of distinction or persistence, in the scenarios of appearance and history of living forms is in a way already inscribed in the structure of the concepts since it presupposes their strict separation. (Bognon-Kuss 2023, 209)

This reappears in postmodernist models as the ‘virtual’ and ‘actual’, so that the (immortal) germ line assumes the role of the ‘virtual’ (Colebrook 2002). Structuralist ideas, by contrast, delineate the organism as a unitary entity. To support this argument, Goodwin, Webster and other structuralist biologists return to earlier French endorsements of biological form, and of transformation (evolution) as involving changes in the structural organisation of the organism.

Gene and Function As the newly named science of biology took hold in Foucault’s modern episteme, function took precedence over structure, the structures that eighteenth-century natural historians had tried to align with a unitary, divine plan, and that Cuvier had reformulated as ‘prerequisites’ to adaptation. A major outcome of this shift was that the objects of study were now described in terms of their adaptation to external conditions based on internal function. Foucault’s thesis in his “Order of Things’ was that Cuvier’s findings were adopted by Darwin as he developed his theory of evolution. However, as suggested by Marjorie Grene and David Depew, Darwin’s theory of adaptation was a concept that was fundamentally “anti-Cuverian” (see above). After Darwin, function became linked to heritable units of variation, rather than Cuvier’s organised ‘structure’. The prioritisation of functional components over organisational properties in life forms has continued to this day in the paradigm of genetic manipulation; gene function, as presented, accounts for the performance of life against the harsh conditions under which it struggles to survive. As such, the gene assumes an almost magical significance, representing the answer to a whole range of agricultural, medical and environmental problems. The new doyen of functionalism in biology is the gene. The gene is difficult to define as it involves more than a code sequence. A typical definition of the gene is: “The fundamental physical and functional unit of heredity, which carries information from one generation to the next; a segment of DNA, comprised of a transcribed region and a regulatory region that makes possible transcription” (Griffiths et al. 2000). By this

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

57

definition, information is transmitted from DNA to polypeptides, which in turn fold into three-dimensional protein structures. But the DNA code would be useless without the intervention of regulators, including transcription factors and other enzymes that edit mRNA transcripts, as well as other kinds of RNA (rRNA and tRNA) and associated enzymes that initiate protein synthesis. Between 50 and 100 proteins may be involved in expression of a single gene. No part of this system works in isolation. Additionally, environmental factors may influence the process. In eukaryotes, the gene sequences are translated into proteins only after the RNA transcript (taken from a DNA template) is edited. Therefore, Evelyn Fox Keller suggests, “We might consider the mature mRNA transcript formed after editing and splicing to be the ‘true’ gene” (Keller 2000). This, she continues, is as if the musical score for an orchestra is rewritten by the players as they are playing it. Her point, in this case, is that post-­ transcriptional processes, not the DNA template, define the ‘gene’. The field of evolution developmental biology, or evo-devo, views the gene as just one, not the only, player in determining development pathways. The phenotype is not reducible to the genotype. Systemic entities are viewed “as mediating the actions of genes on development and evolution” (Gilbert 2003). Genes are also susceptible to environmental influences, leading to developmental plasticity. Whether a turtle embryo develops into a male or female is dependent on the external temperature, for example. Conrad Waddington showed that eye colour in Drosophila (fruit flies) was partly dependent on environmental factors. Embryological research reveals that the position of transplanted cells can determine their fate (Webster and Goodwin 1996). Other studies reveal that homological structures may be more permanent than their associated gene networks (see Chap. 7). More recent work shows that the gene, in many cases, is a secondary factor, and may follow phenotypic changes through a process of genetic accommodation (West-Eberhard 2003). This not only revises the Central Dogma (that genes are determinant of somatic processes) but raises questions about hierarchy. A group of theoretical biologists conducting research in relational biology (see Chap. 5) suggest that bidirectionality (upward and downward causation) in biological systems endorses an anti-­ hierarchical model. However, Mary-Jane West-Eberhard subscribes to the idea of hierarchy in biological processes (it would be difficult to explain such processes without a key component) but views these as essentially

58 

P. MCMAHON

reversible. A gene (even in the same system) may both target and be targeted, depending on external conditions. In any living system, genes are of central importance. Their role cannot be sidelined by a Lamarckian narrative. The problem, raised by structuralists such as Brian Goodwin, is their elevation to the role of an unanswerable director, rather than another component of a system. The discussion here is not to suggest that Darwinian selection is irrelevant, but to place selection in the context of internal and external environmental influences. This is where evo-devo has made major contributions: Just because evolution and development have genetic bases does not mean they have to be reduced to these bases. Only by mistaking developmental biology as reductionistic would we risk the error of getting rid of the genes as explanations. The merging of evolution and development does not mean that we have to throw away our developmental notions of genes or of specificity. (Gilbert 2003)

Selection involves more than the gene: it affects the whole organism in all of its development stages. Nevertheless, the gene (its DNA and RNA structures) lend stability to inheritance. In this sense, genes play an essential role. If genes are followers as well as leaders (West-Eberhard 2005), this only indicates that any hierarchy within living systems is invertible: a gene can be both depending on the context. Systems Biology Entrenches the Reductionist Approach Genetic manipulation now falls under the purview of systems biology. Molecular research reveals that post-transcriptional processes, influenced by factors other than DNA sequences, are just as critical to gene expression systems as the genetic code. This new biology has holistic attributes since the atomistic ideas of the past are now being replaced by an emphasis on interactions and emergent properties. Biological beings are now recognised to be defined by systems, and not by collections of biocomponents that somehow miraculously come together to produce the form of an organism. Biology has moved on from simplistic claims of a one-to-one relationship of genes to adaptive traits. Gene expression is much more complex than this. Holistic properties are emergent from a molecular base, leading to further properties. With the rise of systems biology, supported by high-speed computers, living systems are recognised to have

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

59

holistic properties. However, despite these developments, the genome remains in the ‘director’s chair’ holding to the truism promoted by the Vienna school of logical positivism, that biology can in the end be explained by chemical and physical interactions at the inorganic level (Petersen 2023). The theoretical biologists, Athel Cornish-Bowden and Maria Cardenas, have criticised the current systems approach. Systems biology is not co-­ extensive with the systemic approach of relational biology (see Chap. 5). Current approaches in systems biology, they say, aim for the “integration of knowledge from diverse biological components and data into models of the system as a whole” diverging from the general systems theory (GST) of Bertalanffy (Cornish-Bowden and Cárdenas 2005). The systems theory developed by Bertalanffy is more than a study of the interaction between components as “all of them form parts of a whole, and their presence in the whole can only be understood by considering the needs of the whole” (ibid.). An astonishing quantity of data and computing power is processed, rather than attempts to address the need of the whole system/organism: “Instead of using the view of the whole system as a way to understand its components, it seeks to explain the whole in terms of a vast list of components” (ibid.). The paradigm of modern biology remains reductionist and hierarchical (directed by gene systems), retaining a focus on constituents and their interactions. A system determined by functional needs is contingent upon external conditions. A functional system differs between cultures, or between species, in its organisation of structural parts, but always has a common purpose—to be useful and effective. Therefore, if systems biology has claims to be holistic, this is based on the premise of individual entities working in harmony in organised systems. Its properties are emergent: for example, molecular pathways interact to form the cell, which has new properties. Cells make up tissues, which in turn have their own unique properties. The whole system and its multiple levels operate through homeostatic feedback loops, ending up in the stable organism. A hierarchy is evident, as inner function is responsible for organism structure. Not surprisingly, this model invites criticism, since the organism is so invariant (appearing never to change) despite the continual turnover of its constituent materials, something that Cuvier himself pointed out (Webster and Goodwin 1982). How can a functionalist explanation account for such invariance? We can witness this in our own bodies, which are not the same in terms of their constituent materials, as just a few years ago. Our

60 

P. MCMAHON

skin is being continually replaced, while the bones of our skeletal system are replaced every 8–10 years. Yet form is maintained. How can such stability be achieved by homeostasis? This problem of invariance prompted Hans Jonas to reintroduce the idea of form, important to scholasticism and inspired by Greek philosophy, but more or less dropped in modern science. Systems biology is simply functionalism in a more sophisticated guise, sticking to the main thesis of molecular biology, the Central Dogma put forward by Francis Crick: information flows from genes to proteins and never the reverse. It forms part of the program of technological advancement in the exploitation of nature (so remains firmly within a humanist/ Promethean framework, whether capitalist or Marxist). Functionality Attributed to Material Components Gene function is viewed as the primary level in the hierarchy of the organism. Biology since the modern synthesis (Huxley 1942) proceeds on the reductionist assumption that the answer lies with the physical sciences (chemistry and physics) that define molecular interactions and collection of data obtained through manipulative procedures will in the end provide answers. To the Vienna circle of logical positivism, for example, physics would in the end ‘explain all’; that is, physics explains chemistry, which explains biology, which in turn explains psychology and behaviour (Gatherer 2010). Robert Rosen, who put relational biology on the map, objects to this assumption (see Chap. 5). He proposes in ‘Essays on Life Itself’, that biology should be a science in its own right (and not an outcome of the physical sciences). He suggests that Mendelian analysis is now expressed in terms of molecular fractions, explaining both expression and heredity in one swoop: Today, the problem is considered solved by the identification of everything genetic in an organism with a fraction of molecules called DNAs; all other fractions are epigenetic. Hence the Mendelian phenotypic characters are in there. It is just a question of finding them and identifying them. Moreover, whatever they are, it is supposed that the relation between the genetic fraction and its complement, the epigenetic fraction, is essentially the same as between the program to a machine and the execution of that program as a sequential diachronic process extended in time. (Rosen 2000, 48)

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

61

At the heart of the biology of the modern episteme in Foucault’s account is the quasi-transcendental of life, a mode of knowledge created by the unifying force of function that rules structure. Function remains at the heart of life in this episteme, but becomes a target of the fetishism of commodities. Emerging from this is the gene, a fetishistic object according to Donna Haraway, that is brought to light and transcribed to enable ownership (Haraway 2003). With advances in biology, the gene has become the new El Dorado. Exerting control at this level will generate the greatest profits. To achieve this the gene must be isolated, and brought to the surface in terms of script. The unifying principle is that the gene is now rendered into a concrete object, something that can be commodified and patented. The biology of post-modernity no longer endorses Cuvier’s ideas. The organism constituted by relations has become an irrelevance, replaced by a model of emergent, cybernetic ‘systems’ of material interactions.

The Road to a Historicist Biology Although, as Michel Foucault pointed out, Darwinian theory was inspired by Cuvier’s linkage of the organism to its environment (or conditions of existence), adaptationist ideas were not built upon Cuvier’s coherent organism. As mentioned already, Marjorie Grene and David Depew contend that these ideas were even more anti-Cuverian than anti-Lamarckian. Early French ideas in the discipline of biology (built upon the natural history of the eighteenth century) emphasised structural organisation and immanence in life forms. In their 1982 paper, ‘The origin of species: a structuralist approach’, Gerry Webster and Brian Goodwin, working at the University of Sussex (Fig. 3.1), challenged concepts ruling the biological sciences by returning to earlier ideas that endorse living form (Webster and Goodwin 1982). However, they glossed over divisions within the French natural sciences, counterposing the French ideas of immanent form to models endorsing an archetype or a separate germ line. French biology at the time followed either Cuvier’s functionalism or a unitary model of morphology, but at the heart of each was the notion of ‘form’. Foucault suggests that with his comparative method, Cuvier embodied the move to an historicity that also occurred in other disciplines, notably linguistics and economics. Although linguistic studies began to focus straight away on diachronic change with the epistemic shift from classical to modern knowledge formations—“The new grammar is

62 

P. MCMAHON

Fig. 3.1  Sussex biologists in 1968: (left to right) Jonathan Cooke, Gerry Webster, and Brian Goodwin. (Source: This is Sussex, April 16, 2021)

immediately diachronic” (Foucault 1970)—a diachronicity was not immediate in biology: The same archaeological event was expressed therefore in a partially different fashion in the cases of natural history and language. By separating the characters of the living being or the rules of grammar from the laws of a self-analysing representation, the historicity of life and language was made possible. But, in the sphere of biology, this historicity needed a supplementary history to express the relations of the individual with the environment; in one sense the history of life is exterior to the historicity of the living being; this is why evolutionism is a biological theory, of which the condition of possibility was a biology without evolution—that of Cuvier. (Foucault 1970)

Cuvier remained a fixist, creating the paradox raised earlier (see above). Foucault proposes that by breaking the rational order of eighteenth-­ century natural history, thus challenging the proposed continuity and homology that could not accept large gaps between ‘types’ or taxa, the door was opened to a new account, one based on a new set of relationships. This was a historicity, but one without a history in biology. In linguistics, the comparative approach directly led to diachronic accounts of language (a living language differs between cultures, reflecting those

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

63

cultures, and change with time). But in the natural sciences, a “biology without evolution” was needed before Darwin’s evolution theory could burst onto the scene in the late nineteenth century. What was this ‘historicity’? As Foucault makes clear, it was not based on an idea of change, of transformism, for example. Cuvier rejected this. Rather, it was based on Cuvier’s reformulation of the relation of the internal to the external, of the function of the whole organism in its context. The inorganic environment was no longer another feature to be added on to the end natural history tables (in keeping with the continuity of being), but central to and definitive of life forms. To function the organism depends on its environment, it recycles materials in the environment, it escapes from predators or catches prey and to do this, particular structures and structural arrangements are necessary. External conditions are now fundamentally separate from internal ones, which enable survival through a particular organisation, unique to life. Cuvier insisted that between embranchements these arrangements are not linked; they are non-­ homologous. But in addition, he insisted, they cannot be violated by gradual or incremental changes; they are fixed and unchangeable (for how could they be functional otherwise?). This contradiction could not be maintained—it could only be resolved by introducing a diachronic account and a descent from previous forms with an historical origin, which was achieved by Darwin. Others place less emphasis on Cuvier’s model as a precursor of Darwinism, which could be considered to be “anti-Cuverian” (Grene and Depew 2004, 151). Certainly, as Foucault (1970) points out, Darwin was influenced by Cuvier’s ideas on the adaptation of life to its conditions of existence, but reformulated these in more atomistic terms to accommodate the possibility of change (Grene and Depew 2004, 212). To Cuvier, the organism embodied the idea of form, a form inspired by Aristotle; that is, one that cannot be separated from function, and one that is fixed. His explanation of the fossil record was one of catastrophic extinction, followed by replacement with another series adapted to the new environmental conditions. Any idea of continual change over time was precluded. Evolutionary theory, developed in the hands of Darwin, rejects this fixist notion. Quite apart from the creationism implied by religious beliefs of the time (held also by Cuvier), the Aristotelian standpoint of Cuvier remained a weakness in his theory. Some account of change was needed— in this sense, Geoffroy, inspired by Lamarck, took his theory further than

64 

P. MCMAHON

Cuvier. But it was Darwin, recognising the implications of a ‘historicity’ in Cuvier’s ideas, who could turn evolution into a historical account.

Bringing Back the Formalist Ideas of Early French Zoologists Brian Goodwin and Gerry Webster attempted to outline a structuralist (or formalist) perspective of biology in their 1982 paper. They return to early biological ideas, particularly those developed by French zoologists in the early nineteenth century. They remain firmly wedded to a formalist model of the organism, with Georges Cuvier, Geoffroy Saint-Hilaire and Immanuel Kant as icons of early biology, only to be forgotten with the development of a mechanistic biology. With other biological structuralists, they have levelled heavy criticism towards genetic technologies. This stems directly from their rejection of the Weismann barrier, and the Central Dogma developed by Francis Crick and his colleagues: DNA makes RNA makes protein, namely, that biological information flows one way from the genetic code to its ‘products’ the proteins and somatic bodies. In structuralist terms, the organism is based on relations between its component elements and not on the properties of a particular plan, whether genetic or an archetypal, original form. Now, the organism no longer has its own reality: “The organism as a real entity, existing in its own right, has virtually no place in contemporary biological theory” (ibid., 16). In their paper, Webster and Goodwin side with the materialist tendencies of French thought that prevailed towards the end of the eighteenth-­ century Enlightenment, and are highly critical of the influence of post-Kantian German idealism on modern biology. Movements such as Naturphilosophie challenged the limits Kant placed on access to knowledge of the object, or the ‘thing-in-itself’. In fact, in their paper Webster and Goodwin group post-critical German idealism and British empiricism together, juxtaposing them against the formalist/materialist French biology of the early nineteenth century. The focus of their critique is the emergence of the genetic system in neo-Darwinism as a ‘central directing agency’, no different in its basic premise from the ‘archetypal idea’ that German biologists believed underlay all life forms. Webster and Goodwin counterpose these notions to formalist ideas in early French zoology, including unity of composition, law of balance, functional anatomy, and philosophical anatomy.

3  SETTING THE STAGE FOR EVOLUTIONARY THEORY 

65

To Webster and Goodwin, formalist values bound together the early biological thought of Cuvier and Geoffroy, while their differences (Cuvier’s prioritisation of function over structure in contrast to Geoffroy) were secondary. Also glossed over was the common ground between Geoffroy and German biologists such as Lorenz Oken (1779–1851) with whom he was in correspondence. Unity of type or composition had common features with the bauplan (the deistic plan of life on which all life forms are built) and this forged a bond between French ideas on homology and German idealism. Nevertheless, Geoffroy stuck to material premises in accounting for homologous forms, rather than a mystical archetype. Additionally, while Goethe’s ‘leaf’ has remained stuck onto his legacy as the quintessential archetype, Rolf Sattler suggests the archetypal structure was much less of a feature in his morphogenesis than is often reported—he was more concerned with how the organism develops as a holistic entity. It seems evident that Webster and Goodwin prioritised the unifying themes in early nineteenth-century French zoology over the strong divisions between functional and philosophical anatomy models that became more entrenched in the 1820s leading to the Great Debate. They link the two antagonist groups through common ideas of form and structure: it is less relevant to them whether form is functionally derived (Cuvier) or a result of the unity of composition of animal life (Geoffroy). The key premise in French biology was that some unity was accepted (for Cuvier this was within his embranchements, while Geoffroy and his disciples, believed it could be demonstrated between embranchements). The corollary to this position was that internal elements could not be considered in isolation from their arrangement in the whole system. However, from this point on major differences arise: even though Cuvier was presented as the exemplary formalist by Webster and Goodwin, it was Geoffroy’s idea of a structural plan that seemed to provide viable explanations of transformation (or progressive change with time).

References Bognon-Kuss, Cecelia. 2023. Metabolism in crisis? A New interplay between physiology and ecology. In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T.  Wolfe, 193–216. Switzerland: Springer. Colebrook, Claire. 2002. Gilles Deleuze. New York: Routledge.

66 

P. MCMAHON

Cornish-Bowden, Athel, and M. Cárdenas. 2005. Systems biology may work when we learn to understand the parts in terms of the whole. Biochemical Society Transactions 33: 516–519 . https://doi.org/10.1042/BST0330516. Darwin, Charles. 1859. On the origin of species. 1st ed.: Harvard University Press. Foucault, Michel. 1970. The order of things: an archaeology of the human sciences. New York: Pantheon Books. Gatherer, D. 2010. So what do we really mean when we say that systems biology is holistic? BMC Systems Biology 4 (1): 1–12. https://doi.org/10.1186/ 1752-0509-4-22. Gilbert, Scott F. 2003. Evo-Devo, Devo-Evo, and Devgen-Popgen. Biology and Philosophy 18: 347–352. Grene, Marjorie, and David J. Depew. 2004. The philosophy of biology: An episodic history. New York: Cambridge University Press. Griffiths, Anthony J.F., Jeffrey H. Miller, David T. Suzuki, Richard C. Lewontin, and William M.  Gelbart. 2000. An introduction to genetic analysis. 7th ed. W.H. Freeman & Company. Haraway, Donna Jeanne. 2003. The Haraway reader. New York: Routledge. Huxley, Julian. 1942. Evolution. The modern synthesis. The Modern Synthesis: Evolution. Keller, Evelyn Fox. 2000. Century of the gene. Cambridge, MA: Harvard University Press. Petersen, Eric L. 2023. A ‘Fourth Wave’ of vitalism in the mid-20th century? In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T. Wolfe, 173–192. Switzerland: Springer. Rosen, R. 2000. Essays on life itself. Edited by Timothy F.H.  Allen and David W.  Roberts. Complexity in Ecological Systems Series. New  York: Columbia University Press. Webster, G., and B.C.  Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-­1750(82)91390-­2. ———. 1996. Form and transformation: Generative and relational principles in biology. Cambridge UK; New York: Cambridge University Press. West-Eberhard, Mary Jane. 2003. Developmental plasticity and evolution. Oxford University Press. ———. 2005. Developmental plasticity and the origin of species differences. Proceedings of the National Academy of Sciences of the United States of America 102 (Suppl 1): 6543–6549. https://doi.org/10.1073/pnas.0501844102.

CHAPTER 4

Adaptationism and the Author

Adaptationism: A Functionalist Paradigm Rather than turning to interactions between whole organisms in an ecological context, the basis of agroecology, the paradigm of genetic manipulation is based on the story of adaptationism. The gene is viewed as the chief adaptive component: it ‘maps’ the environment. As the adaptationist paradigm states: change the internal genetic makeup and we can improve the capability of the organism to cope with the external environment. Genetic manipulation fits well with (and indeed is a product of) the functionalist paradigm that informs biology currently: its aim is to modify internal function directly (that is by using direct DNA insertions rather than traditional crossing) to improve performance in the face of external pressures and problems. However, it should be remembered that the gene works within a system. As Donna Haraway has pointed out, the gene is not an enclosed, isolated object—it belongs in a set of relations with associated proteins, other metabolites and cellular processes. The gene is the subject of commodity fetishism according to Haraway, and “seems to be the source of values” (Haraway 1997, 143). The rise of systems biology has led to a more contextualised view, although, as suggested in Chap. 3, this development remains based firmly on a reductionist premise, explaining organisation by an upward causation. It endorses the main principle emerging from the molecular discoveries of the 1950s and 60s: the Central Dogma, DNA makes RNA makes protein. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_4

67

68 

P. MCMAHON

Systems biology has grown rapidly in recent years, made possible by the increase in computing power due to the massive amount of data generated: Systems biology, thanks to high-throughput technologies, is helping modern geneticists to perform, on a larger scale, the same type of studies performed on a small scale using classical linkage studies. Using powerful computational methods, it is now possible to integrate information maps deriving from different biological fields, thus helping in identifying candidate genes and understanding their functions. (Cuccato et al. 2009)

This is valuable research, but its ultimate goals are to find new ways to manipulate organisms. Here are some of the end goals of systems biology according to the US Department of Energy: sequester excess carbon, produce and use energy and clean up the environment. All this could be achieved by manipulating microorganisms. Under the new systems approach, the assumption underlying the ethos of genetic manipulation is not so much that the gene works in isolation (although descriptions such as ‘The Selfish Gene’ may give that impression) but that it (and the systems it directs) are the organism’s internal solution to external conditions. The master systems directed by the gene hold the ‘key’ to a whole swathe of agricultural and environmental problems, rather than the whole organism in the context of its ecological relations. As such, the environmental responsiveness of the organism is downplayed by enterprises of genetic manipulation—in this discourse, answers are to be found in additive functional attributes, rather than the whole organism or contextual system. Molecular biologists engaged in genetic engineering base their applications on one of two suppositions: the first is that knowledge of the transferred gene’s functionality in its original genome will enable a prediction of phenotypic performance in the transformed organism, and the second is that the outcome of transformation cannot be exactly predicted but that, on the contrary, transferring a gene is an attempt to impose a change to the system, followed by evaluation of its efficacy. Most biologists these days do not subscribe to the one gene-one trait dogma that dominated the modern synthesis for so long, as it becomes increasingly evident that post-­ transcriptional processes and whole systems of gene expression define traits, not the transcription of a single or few genes. It follows, therefore, that genetic manipulation is accepted to be based on the second option, a heuristic process, or a process of trial and error.

4  ADAPTATIONISM AND THE AUTHOR 

69

Also well known is that any genetic insertion or deletion will have effects on other parts of the system, or in other words, side-effects. Companies and government laboratories that promote genetic manipulation go to great lengths to minimise this aspect of the technology in their reports, while in the meantime detractors dig up evidence that GMOs could be detrimental to health due to the unpredictable effects of directly altering genomes. They may become targets themselves. A famous case is the persecution of Arpad Pusztai, a senior researcher at the Rowett Institute in Scotland, whose team fed GMO potatoes to rats, resulting in increased thickness in the epithelium of the gut wall. The potatoes were modified for resistance to insect pests with a gene that expresses lectin (found naturally in snowdrops). After he reported his results he was promptly sacked (Guardian 1999). Later his work was peer-reviewed and found acceptable for publication. A functionalism that underlies adaptationism and endorses the gene as the primary locus of change (leading to new structures in evolution) is consistent with current applications of biology, especially genetic manipulation technologies. Therefore, alternatives to the adaptationist explanation of life forms may also, by default, be critiques of the assumptions that legitimise the focus on these technologies as the main purpose, the perceived end-point, of biological applications. That is, while the technologies themselves could be useful (like most technologies), the unquestioned discourse that genetic technologies must become the primary aim of the life sciences, their raison d’être, becomes suspect in the light of alternatives to a functionalist biology, discussed in later chapters. Alternative ideas have been raised that question adaptationism as the primary evolutionary mechanism—although these alternatives generally also include adaptation in their models, in some form or another, for example, within populations. In this chapter, the one-way relation of gene to adaptive structures in the adaptationist account is raised in comparison with literary theories that emphasise the author as the locus of meaning.

The Adaptationist Account Following on from the disagreements among early French biologists, outlined earlier, debates continue today on whether biological diversity (the plethora of species) is built upon common structures (unity of type) which are honed or modified by natural selection, or whether all life forms have evolved in response to their conditions of existence:

70 

P. MCMAHON

Shall we regard the plan of high-level taxonomic order as primary, with local adaptation viewed as minor wrinkles (often confusing) upon an abstract majesty? Or do the local adaptations build the entire system from the bottom up? (Gould 2002, 252)

Despite alternative views of biological form and structure receiving more attention as some of the premises of the modern synthesis are questioned, the adaptationist (also called functionalist) version is favoured by (post)modern biologists. According to this position, “Biological forms [are] the result of internal chance and external necessity” (Goodwin 1990). As illustrated by the moth shown in Fig. 4.1, remarkable adaptations of mimicry and camouflage are found among insects. The adaptationist position is that gradual cumulative genetic changes enhance fitness. In the case of mimicry, each selected genetic (heritable) mutation enables more progeny to be produced since they are more difficult to locate by predators. In the moth species, each individual carrying that improved

Fig. 4.1  The moth Gastroparcha padale mimics a dead leaf. (Vaishak Kallore, Creative Commons)

4  ADAPTATIONISM AND THE AUTHOR 

71

gene (that makes the wing more ‘leaf-like’) produces more progeny, so that the gene spreads through the population: “As far as form is concerned, the organism is thought of as an aggregate of parts each of which stands in functional relation to the external environment (and is therefore subject to natural selection) and in constitutive relation to a set of determinants in the germ plasm” (Webster and Goodwin 1982, 29). In addition, genetic recombination (found in sexually reproducing organisms) contributes to variation. The systems approach to gene expression has revealed the role of apparently unrelated (modifier) genes in determining traits. Therefore, a beneficial character of mimicry might be initiated by a mutation in a regulatory gene, affecting a major pathway of gene expression, and a correspondingly large phenotypic change. Following such a mutation, gradual changes then account for the fine-tuned evolution of the character. These favourable changes accumulate over generations increasing fitness more-and-more, and over time the lepidopteran wing develops the remarkable adaptation of mimicry. “Living organization is the result of cumulative selection” (Dawkins 1991, 45). This explains how chance alterations in the code can result in astronomically unlikely complex forms, such as the eye, given the luxury of geological time (time-scales that we find difficult to grasp). “To tame chance”, Richard Dawkins contends “means to break down the very improbable into less improbable small components arranged in series” (ibid., 317). And, as Dawkins says, natural selection is anything but random. Yet the additive nature of the accumulation of singular changes in the genetic code is very atomistic—half an eye might be better than no eye if it detects predators (albeit in a blurry fashion), as Dawkins suggests, but unless the trait is straightforward (and improves fitness on the basis of single genetic mutations), it is difficult to envisage mutations (which are usually deleterious) as conferring an additive advantage that can be built up into complex systems. Furthermore, a one-one gene-phenotypic character relation is a myth since “a single gene may affect a wide range of traits and, conversely, many separate genes often combine to produce a single trait” (Capra 1997, 220). Therefore, Fritjof Capra writes, “It is thus quite mysterious how complex structures, like an eye or a flower, could have evolved through successive mutations of individual genes”.

72 

P. MCMAHON

Darwin’s Synthesis, Adaptation, and the Gene In the modern episteme of biology, Cuvier’s (fixed) organism was transformed into Darwin’s account of evolution by adaptive change (Foucault 1970). To achieve this Darwin adopted two principles from earlier thought: (1) Despite Cuvier’s denial of transformism, change did occur Darwin observed—this was supported by the French rationalist ideas of Bonnet, Lamarck and Geoffroy, by the ideas of his own grandfather, Erasmus, and those of Charles Lyell on geology, and by his own observations of selection in domestic farm and garden crosses. The question for Darwin was: How was change and speciation effected? (2) Organisms were admirably suited to their environmental conditions, demonstrated by Cuvier. This insight was supported by Darwin’s own studies, including his observations during the voyage of the Beagle. Darwin proposed that change occurred gradually over many generations by the cumulative, incremental addition of characteristics affected by variability. These variations were heritable, and only selected if they conferred a survival, therefore, reproductive advantage. Darwin could not account for the mechanism of variation, which was to be provided much later in the Modern Synthesis, when his theory was merged with genetic discoveries. In the age of the molecular ‘gene’ instigated in the 1950s, singular characters, the expression products of DNA, were envisaged to be the drivers of phenotype. Now they could be materially described under the banner ‘DNA makes RNA makes protein’. The genetic rule over organism function fits in well with the adaptationist view, which has been taken to an extreme, proposing that life forms can assume any form, depending on requirements, putty in the hands of adaptation. All adaptive structures can be explained in this way, according to adaptationists. Genetic action translates directly into phenotypic change. However, the Austrian biologist Rupert Riedl (1925–2005) saw the modern synthesis as limited in its ability to explain the bauplan or morphological arrangements, which he proposed arose through other mechanisms of macroevolution. Others also raise objections to an adaptationist explanation for all biological structures. Muller and Newman (2003, 7), for example, regard this explanation as inadequate: neo-­ Darwinism “completely avoids the origin of phenotypic traits and organismic form. In other words, neo-Darwinism has no theory of the generative”.

4  ADAPTATIONISM AND THE AUTHOR 

73

The adaptationist view proposes function or content constitutes the adaptive possibility of survival. This has been extended to genetic function: if we can decode gene function, we can tap into this information to provide useful characters in genetic transformations. Meaning or content in biology now rests with a new form of life, the gene. The target genes of transformation could be used to express useful characters in different organisms. For, notably, it is not the structure of the organism that is of interest to (post)modern biology, but its meaningful content or function. August Weismann (1834–1914) separated the organism into two distinct parts, “one of which was mortal and transient (the body) while the other was potentially immortal, the transmitter of hereditary instructions (the chromosomes in germ cells)” (Goodwin 1994, 27). The informational ‘program’ is located in the germ cells according to the Weismannian thesis. The germline, in turn, determines the fate of the bodily cells (the soma) which constitute the structural arrangement (the anatomy) of the organism. After Weismann, Brian Goodwin contends “organisms became nothing but the vehicles of genes” (ibid., 27). Heritability remains firmly the province of the germline. The (resistant) structures in biology are overshadowed by the meaning or content of life. With the development of a modern biology, the pendulum swung away from notions such as immutable form to the adaptive potential inherent to life. Now this potential lies with the gene; the gene contained by the Weismann barrier is the final determinant of all somatic structure, reiterated decades later by the Central Dogma of molecular biology (Jablonka and Lamb 1995, 49). Genes explain all living phenomena, even behaviour according to proponents of sociobiology and evolutionary psychology, such as E. O. Wilson (Prindle 2009). Fredric Jameson regards sociobiology as based on myth and ideology, attempting to “link complex societies with the simplest of biological urges”. When a myth is deployed as positive, he maintains, it is “for the most part ideological” (Jameson 1991, 129–130). Stephen Jay Gould remarks that the program of sociobiology “rests upon the view that natural selection is a virtually omnipotent architect, constructing organisms part by part as best solutions to problems of life in  local environments” (Gould 1994). Therefore, Gould maintains, adaptationism leads to genetic determinism. This has strong political implications, as adaptationism endorses the status quo becoming aligned with conservative thought (Prindle 2009). Roger Lewontin and Gould’s 1993 paper ‘The Spandrels of San Marco and the Panglossian Paradigm: A Critique of the

74 

P. MCMAHON

Adaptationist Programme’, was seminal in launching a critique on an adaptationist program they viewed as taken to an extreme. Gould himself did not reject adaptationism as a powerful explanation of how organisms change in relation to the environment. Nevertheless, the criticism unleashed by Lewontin and Gould had as its motive a desire to rein in the more extreme implications of adaptationism, and the discipline that championed it, sociobiology (Prindle 2009). Under this discipline, every gene was seen to have an adaptive function: uncover this and then the gene becomes a predictor of every possible trait, from vision to behaviour. Culture has little to contribute to behaviour in this scenario: “Predictionism and adaptationism, by assuming that genes hold culture on a short leash, denied to humans the ability to construct their own patterns of behaviour, and therefore left them subject to historically dominant modes of unjust domination” (Prindle 2009). Nevertheless, those supporting the sociobiology program deny having a political agenda, regarding their discipline as remaining solely within the confines of science, something that David Prindle on the whole accepts. Proponents of adaptationism propose that the most important driver of evolution is the action of natural selection on mutations in the genetic apparatus. In other words, a new feature is introduced to an individual organism that may, or may not, increase its fitness and ability to pass on the change. By the adaptationist/Darwinist narrative, chance mutations (agents of variation) are selected under particular environmental conditions. If fitness conferred by the changed heritable factor (the altered code or other mutation) to particular progeny is increased, then the progeny will be more likely to survive and reproduce. The modern synthesis attributes variation to a change in the genetic code due to chance mutation events (acknowledging caveats such as mutation hotspots in DNA or the propensity for some proteins to change more rapidly than others) and to genetic recombination. Any chance alterations (mutations) that do not improve an organism’s adaptation to changes in the environment die out, while those that improve reproductive success are passed on to following generations. Over long periods of time and through isolation mechanisms such geographical separation, cumulative variation could account for speciation (also known as macroevolution). This has been the dominant thesis, informing biology to the present day. By this theory, we would expect a gene for every adaptation and might predict that different phyla would not have genes in common—phyla, identified first by Cuvier, are taxonomic groups with practically nothing in

4  ADAPTATIONISM AND THE AUTHOR 

75

common structurally (Chap. 2). This was indeed the prediction of evolutionary scientists, such as Ernst Mayr. This has now been disproved— genetic homology between organisms (meaning the same gene occurs in different species) is now the norm in biology (Rose and Oakley 2007). Most genes have homologues (with slight modifications) in other species and more recently deep homology has been demonstrated by the discovery of genes common to insects and mammals (ibid.). Rose and Oakley (2007) propose that as evolutionary theory and molecular biology are combined under the banner of genomics, the modern synthesis is crumbling. But rather than raising questions regarding genetic manipulation, the new biology seems to reinforce the Baconian paradigm, that manipulating nature is the primary responsibility of science. Structuralist concepts pose questions that undermine this paradigm by returning to the identity of the organism as a unit in its own right, rather than the outcome of gene expression. However, structuralists, like Goodwin, recognise that the gene is still central to any concept of the organism, so long as it is placed within an overall system. Similarly, applications such as quantitative breeding in agriculture focus on genetic effects in developing new varieties (Chap. 10)—but these are integrated with environmental and other (phenotypic) sources of variation. Early Formalist Ideas in Biology Gerry Webster and Brian Goodwin in their 1982 paper counterpose the formalist biology of early nineteenth-century France, to British empiricism, on the one hand, and German idealist endorsements of an archetypal form, on the other. They identify in these early formalist ideas an immanence to life—the organism has immanent properties which are not defined by a ‘central directing agency’. Weismann’s germline and the archetypal form of German idealism both fit the description of such an agency, in which a separation of the material and idea (or an informational ‘director’) are endorsed. Thus formalist models are counterposed to the German archetype and the Weismannian germline/soma separation, which August Weismann proposed to be an affirmation of Darwinian theory. Webster and Goodwin developed structuralist concepts (Webster and Goodwin 1982, 1996), finding common ground with some of the early French concepts in the developing science of biology. However, early French zoologists were far from being in a state of harmonious agreement. Biological issues were hotly debated at the time

76 

P. MCMAHON

causing major divergences between factions, and clear rifts, both professional and personal (Appel 1987). The main point of contention was the primacy of function and structure in defining the biological organism. As outlined in Chap. 2, this difference of perspective led to the debate of 1830 at the Academie des Sciences between Cuvier and Geoffroy. Neither these disagreements (that became particularly acrimonious in the 1820s) nor the debate itself arises in Webster and Goodwin’s discussion. They conflate these differences under the banner of a formalist/structuralist approach to biology. They turn to the transformist ideas of Lamarck and Geoffroy as alternatives to Darwin’s theory of evolution, yet Cuvier, who developed an Aristotelian, functionalist idea of form, strongly rejected ideas that species change incrementally over time. Underlying the morphological formalism of Geoffroy is a strong belief in vital material explanations which prevailed in France at the time. He spoke of fluids and forces—not unlike Lamarck. Cuvier was contemptuous of all such ‘speculation’ and always emphasised the importance of ‘positive facts’ (Appel 1987). In addition, the extreme materialism in vogue at the time clashed with Cuvier’s ideas on functional adaptation. Ideas of spontaneous generation were maintained, for example, by Lamarck. In other words, life could be explained in the end by vital material interactions, and not by its own peculiar properties. This clashed with Cuvier, who developed a model of a functional structure that set apart life from non-life. Life was clearly demarcated from its external milieu due to its internal organisation. Here Cuvier endorsed the Kantian differentiation of the organism from the machine—in the organism, every ‘part’ is involved in the production of other parts. These ideas have returned in current thought as ‘relational biology’ (see Chap. 5), a theoretical-mathematical account of systems developed by Nicolas Rashevsky and his student Robert Rosen: life has fundamental properties that cannot be explained in terms of physics and chemistry, a theme developed by Rosen in his ‘Essays on Life Itself’. Toby Appel’s (1987) account makes it clear that Cuvier at first did not oppose the idea of a unified theme that ran through all animal forms, but also that he aimed to base this on explanations that related life forms to their conditions of existence. Later his own detailed studies led him to propose that God had created four branches (plans) of animal life that were unified by homological resemblances within, but not between, the animal groups (or phyla in modern terms). What he refused to accept was that (1) the four embranchements had common structural arrangements,

4  ADAPTATIONISM AND THE AUTHOR 

77

that is, were homological (or analogical, the term used at that time), and (2) that the same structure could be transformed to assume different functions. So, as an example, the pentadactyl limb in its various manifestations (adaptations to climb, to walk, to swim) comes under the great function of locomotion. But proposals that structures associated with circulation, for example, could assume other functions, whether in a transcendental biology or through transformism, were rejected outright by Cuvier, leading to divisive disagreements within the scientific community. Two twentieth-century schools of literary formalism followed similar paths to formalist biology. Is the organism (text) meaningful and complete in itself? In other words, is it a final cause? Or does it obey structuralist (non-teleological) rules of transformation? Is a work of literature (worth such a label) a holistic entity, each part supporting a meaningful whole? Or is it a changing entity as constitutive devices struggle for dominance with no particular end-point? And contra formalist ideas, is it a product of its environment (social, psychological, biographical and historical factors)? In a similar way to contrasting accounts of the biological organism, literature has been proposed to be either a product of its environment (and external influences) or the result of an engagement with its own form, its internal devices and language. At the time of the first world war, Russian theorists developed formalist models of literature, later to be combined with structuralist linguistics in the Prague Circle (Chaps. 5 and 6). Further west, others developed an Anglo-American formalism, New Criticism, in which meaning was linked to the whole in text, especially poems: text could only be meaningful as a whole due to internal relations between its parts.

The Author, Romanticism, and German Archetypes Just as for the discipline of biology, literary criticism comes up against problems such as the autonomy of a text, and whether literature evolves due to internal processes or in response to the influence of external factors. Formalist views, emerging with modernism in the twentieth century, criticised the notion of literature as a product of its social environment or authorial intention. As such the Russian formalists were accused of an elitist ignorance of the revolutionary principles of the Soviet socialists (see Chap. 1). Another school, New Criticism, which emerged independently in the United States, was also highly critical of searching for the meaning of a literary text in its external historical and social context, or in the mind of the author or reader.

78 

P. MCMAHON

New Criticism New Criticism was one of the pre-eminent theories of literature in the early to mid-twentieth century. Associated with a degree of conservatism originating in the American south, and becoming established at Yale, this school opposed itself to more left-wing approaches to criticism, particularly among the immigrant community of New  York. The New Critics rejected the prevailing assumption that the meaning of the text could be attributed to an author or external historico-social factors; meaning was found in the text itself without any recourse to authorial intention or reader-input. Typical of formalism, the focus was internal, rather than on external influencing factors. In particular, the inseparability of form and function, as proposed by Aristotle, was endorsed with the idea of the poem as a created form, not an imitation of the eternal forms. To Aristotle, poetry was an attempt to create form as it should be (Fry 2012, 57). This separates Aristotle’s notion of form from that of Plato—that is, eternal forms behind actual manifestations, which are a mere shadowy imprint in comparison. The New Criticism revealed that the form of a poem constituted  its meaning. Therefore, content was not considered separately from form. The New Critics spoke of the poem (not ‘poetry’, the effusion of expression of the Romantics) seeing it as a self-contained object, one with vital functionality—that either works or doesn’t work: The poem was that which could not be paraphrased, expressed in any language other than itself: each of its parts was folded in on the others in a complex organic unity which it would be a kind of blasphemy to violate. (Eagleton 1983, 47)

A complete text like a poem is an organised object, adrift and free from its author and reader. The New Critics wished to make literary analysis more objective, seeing methods that appeal to the author as being too subjective. William Wimsatt (1907–1975), for example, viewed the text as a stand-alone object, left adrift with no connection to its historical context. The text is detached from the author and “goes about the world” belonging to the public (Wimsatt and Beardsley 1946, 5). The text should therefore be analysed objectively, without recourse to possible influences outside of that text, particularly the author’s biography or fashionable ideas of historical progress. In ‘The intentional fallacy’ Wimsatt and

4  ADAPTATIONISM AND THE AUTHOR 

79

Monroe Beardsley (1915–1985) reject “the author’s authority over the text with the concept of meaning as prior to expression, which confines the text for all time to a single and univocal reading located somewhere other than the printed pages” (Belsey 2002, 14). A second publication, ‘The affective fallacy’, similarly criticised the role of the reader (his/her background, beliefs, social environment) in attributing meaning to a literary work (Chap. 9). There is a passive aspect to the New Critics idea of the text; it ‘writes itself’ and is produced as a completed object, rather than being a work in progress. In contrast Yuri Tynianov, the Russian formalist, saw all literary pieces as a drafts, or works in progress (Khitrova 2019). Literary formalists, in Europe and the United States, eschewed external factors as being important in the development of meaning: the Romantic idea of an expression of an artistic idea in a poem or painting and the reader’s response as contributing meaning to a work of art. Formal devices and how they were arranged were viewed as immanent, and independent of external influences. This did not mean literary developments could not be modified by external influences. Socialist realism with its political backing and persecution of formalists in Russia who did not support the ‘cause’ is a case in point. Works were censored according to the socialist ideal. If the Russian formalist movement itself was short-lived its death was not a ‘natural’ one, according to Daria Khitrova (Chap. 5). Nevertheless, even as context was increasingly acknowledged to be an important feature of literary development in 1920s Russia, and especially later in Prague, literariness was still viewed as internal to literary systems, not an environmental product. Similarly, to the New Critics a text was regarded as an object for analysis on its own terms, with no recourse to external context. Common themes occur between the ideas of the early French biologists and literary formalists, the Russian schools in particular. Literary formalists prioritised the internal features of a text, including its devices and the production of meaning by relations among elements within a text. Similar proposals arose from French rationalist biology. The properties of the text or the organism, their ‘meaning’, are internal yet they are functional, because of their internal organisation. These internal properties are relational: it is their relations in the context of the whole (text or organism) that make the whole ‘work’. As such no individual device or element can account for function: that is only achieved by the whole. In agreement with the New Critics, Russian formalists downplayed factors outside of the text itself. Yet they rejected the idea that a text had no contextuality at all. As Yuri Tynianov pointed out, behind any text there is

80 

P. MCMAHON

an author. New Critics (unlike their Russian counterparts) were not interested in structural commonalities, but how a text, that text, functioned. Tynianov remarked how they claimed a text needs no external reference to interpret, yet can always be seen to rely on some kind of context for interpretation. It would be impossible for a text to be a purely stand-alone object, since it always draws upon its contextual environment, even unconsciously. Russian formalist positions were more open to contextuality. Another criticism of the New Critics, levelled by T.S. Eliot (although he was an inspiration for the school) was that meaning could be interpreted in dozens of ways or more, that I.  A. Richards (1893–1979) called ‘an endless swarm of lively rabbits’ (Easthope 1991, 19). Nevertheless, the New Critics, Russian literary theorists and biological formalists had common ground, as they attempted to explain a text or organism, to understand its meaning, by its immanent properties, and not by an external idea, a genetic program or by a dialectic relation with the environment. The ‘Author’ in Biology Are there parallels to literary criticism to be found in biology? The point raised by Ernst Cassirer in his final New York lecture (Cassirer 1945) is that linguistics and biology are characterised by common structuralist principles. Jean Piaget in his 1971 publication, ‘Structuralism’ also identified common structuralist themes in different areas of study, from mathematics to psychology. Understanding these shared principles might shed light on the current paradigm that dominates biology, and how it differs substantially from structuralist perspectives. Webster and Goodwin, in their 1982 paper, turn to formalist biology in their critique of the adaptationist program, which they equate with Weismann’s separation of the organism into germline and somatic cells, with the direction (the ‘plan of life’) residing firmly with the former. Both German idealism and British empiricism come under fire as invoking a central point of control, an agency, that directs an otherwise passive body (the soma). Genetic continuity is important in any explanation of biological organisation but Mary Jane West-Eberhard (2003, 90) contends, like Webster and Goodwin’s point above, that ‘neo-Weismannian reduction’ has been taken to an extreme: [T]he neo-Weismannian reduction treats continuity of the genetic material as if it means that a complete set of instructions for development is passed

4  ADAPTATIONISM AND THE AUTHOR 

81

intact between generations except for changes that result from meiotic recombination and evolutionary change in gene frequencies due to selection or drift. By this view, phenotypic organization is controlled by the packets of genes passed from parents to offspring; development begins with the new, zygotic genome; the genetic material persists while phenotypes come and go; and phenotypic organization in a new generation begins with genes. This sees the germ cells as immortal, whereas the soma is the mortal part of the body of a multicellular organism.

What is actually passed on intergenerationally is a complex structure, she states, “or a set of cells that springs entirely from the previous generation, [that] is adapted for survival and interaction in the gametic and embryonic environment, and is the active and organized field upon which the zygotic/ offspring genome products and subsequent environments eventually act” (ibid., 91). These structures are temporary, and undergo further changes; “seeds, larvae, tadpoles and embryos” are examples of this. Drawing on evidence obtained by evo-devo studies, this view expresses a critique of the view of the genome as a program, first proposed by Sydney Brenner, or a recipe (Dawkins 1991). Well-known and especially important is the inheritance of maternal somatic material by the embryo of many multi-cellular organisms. In addition, the phenome affects the genome just as much as genes influence phenotype—a two-way interaction is apparent that could be considered non-hierarchical (or, alternatively, a reversible hierarchy). Such ideas emanating from both the evo-devo and the structuralist camps have questioned the adaptationist scenario. In literary terms, it might be said that these biologists reject the role of an author, a locus of meaning that becomes expressed in the final text. Formalist schools of literary criticism reject the proposition that meaning is prior to text. In particular, they target expressive realism in literature, which views art forms (poetry or painting, for example) as expressing the perception, emotions and feelings of an author. As Catherine Belsey puts it, in expressive realism literature reflects the reality of experience as perceived by a gifted individual, who expresses this as a truth. A literary analysis may therefore turn to the author’s biography and the historical background in which s/he wrote to elucidate the meaning of a text. Underplaying the role of the author’s intention, biography and historical backdrop, is consistent with the formalists’ rejection of the notion of literature being the result of a cause, or external influence. However, change or transformation in literature (including its dominant devices, genre) was

82 

P. MCMAHON

never rejected by the Russian formalists (Fry 2012). Literature has its own ‘life’ that is immanent and evolves in unpredictable ways (Chap. 7). Here is an inkling of the formalist critique of the author as the locus of meaning, a meaning expressed in a work of art or literature. Applied to biology, rather than an authorial germline determining a rather passive soma (at least informationally), the soma (the phenotype) initiates change since it already carries alternative modular components (cryptic systems that require an environmental cue to become activated). The phenotype can, therefore, adapt to changing conditions (West-Eberhard 2005). Rather than being determined by genes, it engages with genes in a system. This evo-devo thesis does not preclude genetic variability as important to evolution, or the gene as the most important instrument of heritability, but does undermine genetic determinism. Such determinism is based on the idea that the gene ‘programs’ phenotypic outcomes, just as in the case of literature, the ‘author’ embodies the meaning of the text. Evo-devo biologists move away from this determinist scenario: the organism (the phenotype) has its own autonomous properties, based on relations between its elements, of which genes are one component. Humanist Causes Expressive realism in literature credits the meaning of a text to the author. The author, in turn, is situated in a particular historical and social context. Literature becomes the effect of an external cause. Both neo-liberal and socialist ideologies are wary of an autonomy that removes art forms, including literature, from their sphere of control. Just as realism aims to direct literature in the service of a particular cause, neo-liberal utopian projects currently propose to place the organism and its living processes under control for human ends. The modified organism is, therefore, linked directly to an external (humanist) cause – its purpose is to enhance human life, while its autonomy is systematically removed. The aim to place the organism in the service of an externally driven ‘modification’ goes back to the atomisation of adaptative traits and extreme forms of adaptationism. Critics of adaptationism maintain that living forms are transformed by internal processes not directly linked to functional need. This is the structuralist position that has inherited some of the early ideas in French biology. A literary evolution proposed by the Russian formalists involves an immanent transformation (Chap. 7). And as discussed in Chap. 8, some biological models suggest a reorganisation of the phenotype could be a driver of evolution.

4  ADAPTATIONISM AND THE AUTHOR 

83

Questioning Current Accounts of Evolution Brian Goodwin, Gerry Webster and other structuralist biologists propose that Darwin’s account of evolution, namely that it proceeds through progressive changes necessary for the adaptation of living organisms to their environment, cannot account for the developments in all of the structures found in life forms. In addition to natural selection, other processes are at work. These biologists wish to place biology on a rationalist footing, a return to earlier French conceptions of the organism: This position starts from the proposition that organisms and their life cycles are processes of a particular kind that result in characteristic forms—phyllotactic patterns in plants, tetrapod limbs and eyes—all as natural consequences of the dynamic order of the living state. (Goodwin 1990)

However, on top of these fundamental structures, Darwinian mechanisms are acknowledged—something referred to as the “adaptive mask” by Richard Owen. In other words, diverse modifications through adaptive evolution hide or mask underlying perennial structures. Structures such as the three thoracic segments of insects, the vertebrae of backbones and the pentadactyl limb have remained stable over eons; thus structuralists would divide living form into primal structures and an ‘adaptive’ mask (Denton 2013). The adaptationist explanation is that at some stage conditions and natural selection led to the evolution of all structures—though this is impossible to demonstrate. An example is the putative evolution of the pentadactyl limb by selection. Even if this is accepted, “[we] have to believe that a variable adaptive form became an invariant, non-adaptive form at a particular instant in evolutionary time and was conserved through all the subsequent generations and phylogenetic lines in both fore and hind limbs” (Denton 2013). Alternatives to adaptationism as explanations for evolved biological structures have been raised by Stephen Jay Gould. One example is exaptation—when a trait evolved for one function serves another. Bird feathers, for example, may have evolved for insulation before being applied to flight. Possibly mammary glands were formerly sweat glands before being coopted for the secretion of milk. The sting in some insect groups may have evolved into an ovipositor. Gould suggested characters could be co-­ opted through natural selection, but in some cases non-adaptive characters are co-opted.

84 

P. MCMAHON

Many biologists have moved on from the modern synthesis, instead subscribing to an extended synthesis—but the old division remains between those who emphasise the role of the phenotype (and its integrated systems) and those who seek molecular explanations, albeit more sophisticated ones than a direct gene-phenotype relation. Both groups bring up deep homologies (genes held in common among very different taxonomic groups), but in a different way. Other ideas have been broached such as epigenetic inheritance and developmental plasticity (Jablonka and Lamb 2005). Consistent with the proposals of structuralist biologists, they move away from the idea of the gene as directing the all processes in the soma to a systemic one, where components are determined to an extent by requirements of the system. Structuralist (or mathematical) biology, considered a sub-field of theoretical biology, tries to identify common rules among body plans, rather than a directive functional ‘cause’. Transformation is postulated to occur, but only within the framework of these rules. This branch of theoretical biology seeks mathematically defined explanations to account for living forms. Structuralists propose that evolution (or history) has a much smaller influence on form than in Darwin’s account, or in some accounts even an insignificant role. This has drawn criticism from some neo-Darwinists, who claim hard evidence is lacking (Price 1995). Structures and living forms are investigated by structuralism, which questions the foregrounding of genetic information. Inorganic and physical laws constrain the range of possible adaptive structures that can develop. Structuralist biologists raise extra-genetic phenomena in determining form, such as diffusion gradients, positional criteria determinant of a part’s fate in development and cellular behaviour. They turn to physics for explanations. The gene is a central and important component in the structuralist model (Goodwin 1990), but is placed within a wider physical system. The gene (the DNA molecule and its interactions) is regulated by this system, rather than determining it.

The Organism as a Unit The common (linking) theme that opposes structuralist biology to the historico-empirical approach expressed most strongly in Britain is that the form of the organism operates as a whole. If one part changes, all parts are changed, or are rearranged. A change in arrangements can lead to a changed function. Transformation (whether ontogenetic or phylogenetic)

4  ADAPTATIONISM AND THE AUTHOR 

85

is explained by a rearrangement of pre-existing elements, internal to the system, and not only by the introduction of new elements (such as mutated genes). Further, properties of life are immanent to its organisation. There is no recourse to a directing ‘idea’, a master plan or genetic program to account for living structures. Similarly, the German archetype, the underlying form to all animal forms (expressed also in botany by Goethe’s archetypal plant or leaf) is questioned, as are competing ideas of recapitulation. Von Baer’s explanation for embryological homology (Chap. 7) is also considered inadequate, since his interpretation of recapitulation is based on an underlying archetypal form, a basic form from which variants develop. The formal characteristics of an organism alluded to by Webster and Goodwin are its self-organising properties. They contrast this with directive models, where a passive body depends upon an information-rich agency (such as the germline cells). Both the archetype in early German biology and the genetic program in neo-Darwinism are reminiscent of an author and of a meaning that is prior to the actual organism. When we consider life and text as having common formal properties, it is evident that Webster and Goodwin are criticising the authorial ‘idea’ or plan in developing their formalist position. They point out that both German idealism (including von Baer’s biology) and Weismann’s mechanism invoke a plan that is external to the actual (somatic) organism. In both cases (German idealism and the Weismann barrier) an agent is invoked. They counter both of these biological explanations to Cuvier’s functional relations and Kant’s notion of self-organisation. An organism that functions as a whole does not need an external director, nor an additional vitalist agent (an elan vital) to account for function, for functionality is bound up in the material medium and relations between internal structures. Just as in literary theories, such as New Criticism, intentionality of the agent/author is rejected as an explanation of the meaning inherent to a literary text or organism. Admittedly, the deeper relations that form Foucault’s ‘transcendental objects’ escape the methods of empirical science. Cuvier’s empirical, comparative method reveals a hierarchy: while positivist science readily identifies visible relations such as those between the incisor and the digestive system, at deeper levels the great functions merge into what Foucault (1970) calls “the great mysterious, invisible focal unity”. Yet the crux of Webster and Goodwin’s argument is that life or meaning (the body of the organism or text) is not directed from an external platform, whether the

86 

P. MCMAHON

archetypal ‘idea’ or Weismannian germline, but that the organism works as a whole unit, with its elements operating relationally to produce such unity. Here there are striking similarities to formalist literature. However, problems arise with formalist models, which can be accused of being “timeless, universal and transhistorical” (Belsey 2002, 16). The Prague Linguistics Circle provided a more contextualised account: text could not be considered as absolutely decontextualised from its external environment, could have multiple meanings or functions, and could change with time (see Chap. 6). Neo-Darwinist theory, our current paradigm, states that current biodiversity is principally the result of historical modification in response to environmental pressures or to other processes such as genetic drift. It might be said that the organism has become an ‘expression’, the product of a genetic program, rather than an immanent whole. Principles raised by literary formalist schools strengthen Webster and Goodwin’s argument that modern biology is based on atomistic and reductionist models that neglect the form of the organism. But returning to Cuvier or Kant is not enough to support this—for how can evolution be explained if the functional relations (the ‘form’) of the organism are fixed and cannot be violated without leading to dysfunction. The idea of evolution (transformism) was raised in formalist circles (both for literature and biology). Common overlapping principles occur between these accounts. Chapter 7 raises some of these common principles, and how formalists in both disciplines (literature and biology) approached the issue of change or evolution. Applying formalist principles to biology, so-called content becomes a structuralist product. The gene itself is caught up in a structure. Studies of homologous DNA sequences reveal their mix-and-match construction— sections have been inverted, taken out, moved to other sections of the genome. Where is the gene in this mix? One answer is that there is no clear-cut gene, just an odd mix of DNA sequences, including (in multi-­ cellular organisms or eukaryotes) stretches of apparently useless DNA, commonly called junk DNA. If DNA is considered to be another component in a system of inter-relations, its meaningful contribution only becomes apparent after an editing process. These structures are used for different purposes, that is, for different gene expression systems through editing. The same sequence of DNA (we think of as the gene) can be used to produce different proteins for different functions. Furthermore, in a

4  ADAPTATIONISM AND THE AUTHOR 

87

phylogenetic series (the species shown in a family tree) the same structure can be changed to assume different functions. Genes that are passed down over generations are known as orthologs (Chap. 7). Therefore, the same genetic structure can assume different functions in time (the phylogenetic sequence) and in the same cell. Multi-functionalism among genes is more common than previously predicted. Why (unlike Cuvier’s assertion that this is impossible) do organisms use the same structure to perform different functions (in modern genetic terms, use the same DNA sequence to form different proteins products)? Perhaps the resources needed to form every gene (and its associated systems) from scratch would be prohibitive. In a sense, it is like renovating an old house. Rooms might be merged into larger rooms, extensions added, the plumbing updated—these changes might be more economical than knocking the house down and replacing it with a completely new one. Additionally, the post-transcriptional process (the editing of the DNA copy) enables flexibility as multiple proteins can be made from a single sequence. This accounts for the surprising find in the human genome project—that there are far fewer genes in our bodies than predicted. If a stretch of DNA (we call the gene) can produce multiple proteins, each with a different function, then clearly we need fewer such genes than would be expected on a one gene-one protein basis.

Structural Transformation, Not Malleability In structuralist biology, as spelled out by Webster and Goodwin (1982) turning back to early French biology, life is ruled by formal principles. To Geoffroy, Cuvier’s colleague and rival, comparative anatomy revealed laws of biology, such as the unity of composition. It should be noted that scientists of the time were strongly influenced by Newton and his laws of physics and wanted to find similar ‘laws’ in biology. Cuvier, as discussed in Chap. 2, was also interested in establishing common taxonomic principles: in structural arrangements, within his embranchements, but not between them—his empirical studies found no common grounding in structural arrangements between the four embranchements of life. These early ideas did not consider life to be a plastic material that can be moulded into endless forms of expression. Nevertheless, evolutionary ideas, such as Lamarck’s, were becoming accepted. This was the period when evolution (termed transformism) was presented as a serious

88 

P. MCMAHON

proposition, rejecting the fixism that had dominated natural history since Aristotle. But although transformism had become accepted by some natural scientists by the early 1800s (Cuvier being a significant exception), change was seen as operating within certain limits or constraints, defined by the laws of life. Transformism, especially in Geoffroy’s view, uses the same elements in living structures for different purposes. This biology was formalist as it saw life as a predefined entity, complete with its own laws. Even Lamarck’s theory of evolutionary change was not based on acquired characters (these were used to explain branching of the taxonomic tree) but on a drive to increasing complexity. It is Webster and Goodwin’s contention in their paper that we have lost sight of such principles in biology, which has come under the sway of an extreme adaptationism—a paradigm that readily endorses the promotion of genetic manipulation. Under this paradigm, genetic manipulation appears to be simply a natural extension of a plastic evolution—an evolution where all structures and all changes are possible, so long as they pass the test of survival. This contradicts ideas raised in the early nineteenth century and revived a century later with publications such as D’Arcy Thompson’s ‘On Growth and Form’. In literary formalism, a similar opposition to both authorial intent and interpretation is posed. A structuralist or formal model of the organism does not refer either to an archetype or to an adaptive response to the environment. In literary formalism meaning is seen to lie firmly in the text itself (Chap. 5), excluding authorial intention and with the reader being more-or-less the passive partner. The rejection of the formalist message of these schools (both literary and biological) by the current (post)modern discourse has effectively thrown the baby out with the bathwater. This is unfortunate because, notwithstanding the limitations of formalism, perhaps text is not just putty in the hands of the reader or an expression of meaning residing in the author (as an idea expressed through the vehicle of language). Perhaps the organism has a form (or structure) precedent to its environment. But, rather than returning to a pure formalism, it might be productive to incorporate some formalist ideas into contemporary thought. As outlined in Chap. 6, Roman Jakobson in his aphasia studies demonstrated that language as a structure is metaphoric but made contextual by Saussure’s axis of combination (or metonymy). With Yuri Tynianov, he emphasised that literature is a transformational system, influenced by

4  ADAPTATIONISM AND THE AUTHOR 

89

external factors, yet retaining autonomy (Tynianov and Jakobson 1972). Discussed further in Chap. 9, Lacanian-inspired literary theory (based on Jakobson’s metaphor-metonym polarity) assumes a role for the reader and subject, while retaining structuralist (and Marxist) principles, such as the signifier-signified distinction, and the role of the material signifier in defining conscious meaning. This represents a rejection of extreme structuralist versions (which often resemble an orthogenesis, or a destined evolution), while maintaining a structuralist version of the text (or organism) as autonomous (Chap. 5). By drawing on Saussurean linguistics and ideas of literary change, a more dynamic structuralist model emerges based on the investigations of the Prague Linguistics Circle (see Chaps. 6 and 7). Similarly, by taking up ideas from studies in developmental biology, as well as early French biology (Chap. 2), structuralist biology can endorse more transformative models that can account for change.

References Appel, Toby. 1987. The Cuvier-Geoffroy debate. Oxford University Press. Belsey, Catherine. 2002. Critical practice. 2nd ed. New York: Routledge. Capra, Fritjof. 1997. The web of life: A new synthesis of mind and matter. London: Flamingo. Cassirer, Ernst. 1945. Structuralism in Modern Linguistics. WORD 1: 99–120. Cuccato, G., G.  Della Gatta, and D. di Bernardo. 2009. Systems and synthetic biology: Tackling genetic networks and complex diseases. Heredity 102 (6): 527–532. https://doi.org/10.1038/hdy.2009.18. Dawkins, R. 1991. The blind watchmaker. 2006 ed. London: Penguin. Denton, M.J. 2013. Types: A persistent structuralist challenge to Darwinian pan-­ selectionism. BIO-Complexity 3: 1–18. https://doi.org/10.5048/ BIO-­C.2013.3. Eagleton, Terry. 1983. Literary theory: An introduction. Oxford: Basil Blackwell Publisher. Easthope, Antony. 1991. Literary into cultural studies. London, New  York: Routledge. Foucault, Michel. 1970. The order of things: An archaeology of the human sciences. New York: Pantheon Books. Fry, Paul H. 2012. Theory of literature, the open Yale course series. New Haven and London: Yale University Press. Goodwin, Brian C. 1990. Structuralism in biology. Science Progress (1933–) 47: 227–244.

90 

P. MCMAHON

———. 1994. How the leopard changed its spots: The evolution of complexity. London: Charles Scribner’s Sons. Gould, S.J. 1994. Hen’s teeth and horse’s toes, paperback edition. Norton. ———. 2002. The structure of evolutionary theory. Cambridge, MA: Belknap Press of Harvard University Press. Grene, Marjorie, and David J. Depew. 2004. The philosophy of biology: An episodic history. New York: Cambridge University Press. Guardian, The. 1999. Top researchers back suspended lab whistleblower. The Guardian, February 12. Haraway, Donna. 1997. Modest_Witness@Second_Millennium.FemaleMan_Meets_ OncoMouse: Feminism and technoscience. New York: Routledge. Jablonka, Eva, and Marion Lamb. 1995. Epigenetic inheritance and evolution: The Lamarckian dimension. Oxford, New York: Oxford University Press. ———. 2005. Evolution in four dimensions. Genetic, epigenetic, Behavioral, and symbolic variation in the history of life. Cambridge, MA and London: MIT Press. Jameson, F. 1991. Postmodernism, or, the cultural logic of late capitalism. USA: Duke University Press. Khitrova, Daria. 2019. Introduction to permanent evolution. In Permanent evolution: Selected essays on literature, theory and film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 1–24. Boston: Academic Studies Press. Muller, Gerd, and Stuart Newman, eds. 2003. Origination of organismal form: Beyond the gene in developmental and evolutionary biology. In The Vienna Series in theoretical biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut. Massachusetts: MIT Press. Price, C. 1995. Structurally unsound: A review of how the leopard changed its spots: The evolution of complexity, by Brian Goodwin. Evolution 49 (6): 1298–1302. Prindle, D. F. 2009. Stephen Jay Gould and the politics of evolution. USA: Prometheus Books. Rose, M. R., and Oakley, T. H. 2007. The new biology: Beyond the modern synthesis. Biology Direct 2: 30. https://doi.org/10.1186/1745-6150-2-30. Tynianov, Jurii, and Roman Jakobson. 1972. Problems in the study of language and literature. In The structuralists: From Marx to Levi-Strauss, ed. Richard T. De George, 80–83. Garden City, NY: Anchor Books. Webster, G., and B.C.  Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-­1750(82)91390-­2. ———. 1996. Form and transformation: Generative and relational principles in biology. Cambridge, New York: Cambridge University Press.

4  ADAPTATIONISM AND THE AUTHOR 

91

West-Eberhard, Mary Jane. 2003. Developmental Plasticity and Evolution: Oxford University Press. ———. 2005. Phenotypic accommodation: Adaptive innovation due to developmental plasticity. Journal of Experimental Zoology. Part B, Molecular and Developmental Evolution 304 (6): 610–618. https://doi.org/10.1002/jez. b.21071. Wimsatt, W.K., and Monroe Beardsley. 1946. The intentional fallacy. In The verbal icon.

CHAPTER 5

The Relational Turn

Autonomy in Literature and Biology Literary formalists subscribe to an inherent autonomy of the text and language. The formalists returned literature and language itself to a status of ‘being’, rather than functioning as an unproblematic instrument of communication. Lav Jakubinsky differentiated between practical language “whose purpose is the communication of meaning” and a literature “where a practical aim retreats to the background … and language resources acquire autonomous value”—quoted by Brown (1974). A work of literature such as a poem embraces ambiguity, and attempts to draw attention to itself, to the use of language for its own end, and not a representative tool of outer reality. Language itself (and linguistics) comes into focus, as opposed to culture and social life. Linguistics is “a discipline contiguous with poetics” said Boris Eikhenbaum (ibid.). Literary formalism brings back the focus onto language and literary devices, rejecting the notion of language as a transparent medium for the ‘message’ or that literature is a ‘window onto the world’. Formalist ideas emerged in Russia and America (New Criticism) independently. Both schools placed a focus on the text, rather than background influences. Themes common to biological and literary formalism could lead to a better understanding of the premises of formalism (and structuralism) in general and their critique of functionalist explanations. Formalists and

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_5

93

94 

P. MCMAHON

structuralists downplay the role of the author in conferring meaning (the ‘death of the author’ as Roland Barthes put it). Meaning is made evident by the whole organism or textual object, and not in an external, separate compartment that then ‘expresses’ its intention as the final product. German idealist interpretations of the organism as an ‘idea’ or archetype can be likewise criticised (Chap. 4). Literary formalist theory is based on the premise that form, based on relations, is constitutive of meaning. As Paul Fry says of the Russian formalists: “Part of their platform is everything is form” (2012, 88). Members of the other main branch of literary formalism, the New Critics, view meaning as determined by the contradictions, paradoxes and tensions created by the devices of a poem or work of literature and not by external cultural or historical influences. This isolationist stance raises the question of the contextual situation of literature (and language) in actual events, in society and history—something that was addressed by the Prague linguistic theorists. Biology as a Science in Its Own Right Issues of autonomy are raised by theoretical biologists. Alternative interpretations of life, such as relational biology, go to great lengths to differentiate the science of biology, from physics and chemistry—in effect rejecting the views of the Vienna Circle of logical positivism and the molecular reductionism in biology emerging in the 1960s. At the base of biology, according to these views, is physics. As Richard Dawkins (1991, 15) writes: The biologist's problem is the problem of complexity. The biologist tries to explain the workings, and the coming into existence, of complex things, in terms of simpler things. He can regard his task as done when he has arrived at entities so simple that they can safely be handed over to physicists.

Just as literary formalism regards meaning as embedded in the text itself, not in some other realm, the organism is meaningful due to the immanent organisation of its structural components. This then is the vitalism that formalists have been accused of—it is a vitalism equated with form and not an external elan vital nor an entelechy (see Chap. 1). In the

5  THE RELATIONAL TURN 

95

formalist view of the text, there is no distinction between form and content. Cleanth Brooks, a New Critic, stated that the poem is experience itself, not a description of an experience. The object, Brooks states, is intuited in the medium of language: “language itself …offers the possibility of meaning” (Belsey 2002). In the same way, the organism is its material medium with a formal organisation. One such account was developed by a branch of theoretical biology, known as ‘relational biology’. Relational biologists reject the reductionism of neo-Darwinism and the Vienna school. Ludwig von Bertanlaffy, the founder of general systems theory (GST) and an inspiration for relational biology, and Ernst Cassirer (a critic of reductionism) were close associates of the Vienna circle of logical positivism, engaging in debates on the topic. This indicates this was seen as an unresolved issue among the Vienna scientists. But with the break-through discoveries in molecular biology of the 1940s and 50s, a reductionist view of life became more entrenched. Life was denied any autonomous status. It is exactly the denial of such an autonomy by prominent figures such as Francis Crick that is consistent with, even indistinguishable from, a ‘bioengineering future’. Bioengineering is not simply an additional outcome of extreme reductionism and the denial of an autonomy in life or in biology as a science, but is bound up in its very discourse (Petersen 2023). To Robert Rosen (1934–1998), who more than anyone developed the concepts of relational biology, biology should be given a unique status, and not viewed as an extension of physics or chemistry. In that sense he opposed molecular reductionism (see Chap. 1). Just as literary formalism is concerned with relations internal to the system (or structure), ‘relational biology’ focuses on the systemic aspect of biology and living processes. An autonomy is introduced that separates the science of life from that of its inorganic components. A corollary of this is a rejection of the possibility of simulated models of life using current computer technology, since living forms need no informational input in order to function. A paradox is also introduced as life is dependent on its environment for materials yet retains autonomous form. To these theoretical biologists, life is characterised by relations and not its material constituents, which are turned over continuously. Literary formalism has also insisted on immanent relations, as opposed to external causation, as characterising literariness and its changing form over time.

96 

P. MCMAHON

Formalists Turn the Tables on Functionalism: Literary Formalism in Russia In Russia, around the time of the revolution, two groups of literary theorists, the Moscow Linguistics Circle and, in Saint Petersburg, Opoyaz (Society for the study of poetic language) were formed in response to the psychologism and symbolic art of the time (in which literature was seen to be an expression of an inner unconscious, or inner imagery translated into text). The symbologists were the targets of formalist critiques. The formalists were responding to ideas that literature emerges from unconscious images, with language merely being the vessel of communication. They rejected this, directing their attention to the form of the text itself. They were also, more circumspectly, criticising the idea of literature as the champion of an historical progress, namely the revolutionary change instigated by the proletariats. In the end, this was to prove to be their death knell as under persecution the Russian formalists were forced to retreat. They became targets of Stalin’s totalitarian regime as Soviet realism was threatened by formalism. Yuri Tynianov (or Tynyanov, Fig.  5.1) drew attention, especially for his ideas on literary evolution. Evolutionism and Darwinism were particularly despised by the Soviet socialists as bourgeoisie capitalist forms of propaganda. At one time, Tynianov’s apartment was occupied for three days by the Soviet secret police in the hope of capturing his colleagues who were viewed as dissenters—as they approached, they saw all the lights on, knew something was wrong and turned around, avoiding almost certain arrest. Yuri Tynianov and Boris Eikhenbaum (1886–1959) lost their positions: “Tynianov and Eikhenbaum were fired from the State Institute of Art History for ‘ideological unfitness’; any new public associations (including literary ones) were virtually banned; censorship and attacks in the press grew to a feverish pitch (the very word ‘formalism’ became a slur and has retained that connotation to the present day)” (Khitrova 2019, 6). Others were arrested or escaped, as did Viktor Shklovsky (1893–1984) in 1922, although he returned the next year claiming he was now ‘onside’ with revolutionary principles. Roman Jakobson emigrated to Prague, and later to the United States. These poets and thinkers regarded the formalist approach to be an affirmation of vitality and life in a grey system. The catch cry of defamiliarisation was to make things strange again, reminiscent of Pound’s ‘make it new’. That this dull grey backdrop was a general comment on society, not the Soviet one in particular, made no

5  THE RELATIONAL TURN 

97

Fig. 5.1  Yuri Tynianov (1894–1943) (Creative Commons)

difference—formalists lacked the revolutionary zeal one would hope for in such prestigious academics and poets. The Russian formalists placed the text itself, its language, under the microscope. To the formalists, language was not an unproblematic vehicle of expression, but an end unto itself. Literature was not a window onto the world. Further it was to be considered separately, as much as possible, from its contextual milieu, the social, political factors external to it—literature was therefore given an autonomous status. The role of literature according to Viktor Shklovsky was to defamiliarise, to make things strange again. Hence, literature did change, not by adapting to its environment, but internally as formal devices assumed dominance. This was necessary as when dominant devices in place no longer defamiliarised, reading became an automatic process of ‘knowing’ rather than ‘seeing’. The role of literature was to raise awareness of the act of perception, while ‘familiarity’ was an act of recognition or ‘knowing’. The Russian formalists, as the New Critics, focused on the poem. One

98 

P. MCMAHON

literary device is the use of units and their boundaries, that is, the lines of a poem. An example provided by Sandra Rosengrant (1980) is taken from WH Auden’s lines, “The leader looking over/Into the happy valley”. Pausing the line after the word ‘over’ gives it emphasis and a greater sense of space is provided, whereas in the prose version ‘over’ lacks any semantic value: “[We] involuntarily stress the word ‘over’, making explicit a spatial connotation that had merely been latent”. Similarly, the word ‘dark’ (temynj) is stressed in the lines of a poem by Yakov Polonsky: The river Kura roars, breaking against the dark Precipice in a living wave

By isolating ‘dark’ from the phrase ‘the dark precipice’, “its secondary, emotional colouring is brought out” (ibid.). This raises Tynianov’s idea of secondary attributes (or lexical colouration) carried by words despite their context. The structure of a poem is primary to meaning, but words have their own autonomy, something that was emphasised by the formalist critic, Mikhail Bakhtin. Lexical connotations, rhythmic factors, such as unity and tightness, and other devices are put into practice to add meaning to a poem. The formalists wished to differentiate practical language (as in some kinds of prose) from poetry. Imagery has different functions in the two forms: in prose it can help to clarify and improve apprehension, but in poetry the object is to “reinforce impressions” using devices to ensure “they are not easily apprehended, that they involve ‘estrangement’ (ostranenie)” (ibid.). Literariness (differentiating literature from other texts) and its formal devices (such a rhyme, metre and repetition in poetry) were defined by this ability to defamiliarise. Literariness was viewed as absent in a piece of journalism, which used practical language aimed at transparency and optimising understanding. Shklovsky wrote in his ‘Art as Technique’ (or ‘Art as Device’): The purpose of art is to impart a sensation of things as they are perceived and not as they are known. The technique of art is to make things unfamiliar, to make forms difficult, to increase the length and difficulty of perception, because the process of perception is an aesthetic end in itself and must be prolonged. (Shklovsky 1992)

5  THE RELATIONAL TURN 

99

The Russian formalists tried to define literariness through science, identifying particular devices common to literature (and not found in other texts). Devices such as repetition, meter and rhyme brought focus onto language itself, revealing it as a form, rather than a vessel for transparent communication. Like the biological formalists of the early nineteenth century (Chap. 2) they wished to identify a taxonomy, the common structural features that gave texts their literariness. They excluded practical texts from this group. This was later challenged in postmodernist circles, beginning with Jacques Derrida. Texts cannot be divided into elitist and mundane forms it was insisted—this is a general criticism levelled at modernism, which became associated with elitism. The elitist view of literature and resulting targeting by the Stalin regime meant that the Russian formalists had become largely inactive by 1927. However, to Daria Khitrova the movement died too soon: “Formalism didn’t die of natural causes; on the contrary, its life was cut short just when it was getting started” (Khitrova 2019, 6), Nevertheless, Russian emigres continued the tradition, particularly in the Prague School (see Chap. 6) which was active until the war years. To Shklovsky, a literary work was nothing but the sum of its devices. Literature ‘lays bare’ its own devices. In this sense, literariness is not consistent with realism, since realist literature aims to mask literary devices, giving an impression of ‘transparent’ content (Belsey 2002), and bringing it closer to the practical language of journalism. Art, on the other hand, makes things unfamiliar. Shklovsky provided an example from Tolstoy— rather than naming an event the story describes it, using other words that describe an act (such as flogging). In another Tolstoyan example, the narrator of the story is a horse, lending a new meaning to ‘straight from the horse’s mouth’. From the horse’s point of view, ‘known’ matters to humans such as ownership of a tract of land seemed to make no sense to the narrator since the ‘owner’ had never set foot on it. Such devices must eventually become familiar (or automatic), driving literature to adopt other techniques to avoid ‘habituated perception’. Thus literature is far from static, but changes in sudden leaps (or shifts) from one dominant device to another, or even to another genre. The formalists were inspired by Futurist poets, such as Velimir Khlebinkov and Vladimir Mayakovsky. These poets rejected traditional grammatical forms, introducing neologisms—or playing with suffixes— and treating the material of language in a similar way to cubist painters. Khlebinkov wrote a poem based entirely on permutations of the word,

100 

P. MCMAHON

laughter—‘smekh’ in Russian (Brown 1974). In a backlash against traditional content-based genres, this iconoclastic modernist movement abstracted form from its background. The futurists placed more focus on word-play, disregarding content; their manifesto was titled “A slap in the face of public taste” and called to throw literary canons like Pushkin, Dostoyevsky and Tolstoy “off the steam boat of modernity”. Yet they supported the Russian revolution (since it represented change and a new future, shaking off traditional modes and beliefs). Despite this support (with Mayakovsky distributing socialist literature before the revolution), the ‘revolutionaries’ were wary of them. Mayakovsky (who took his own life) was heavily criticised by RAPP, the Russian Association of Proletarian Writers. As with later criticisms of the formalists, the futurists were seen as concerned with trivial matters, at best a distraction from revolutionary ideals and at times, contravening them. The formalist poets and theoreticians also came under attack by the socialist regime. They were criticised by Marxist theorists, including Leon Trotsky and Valentin Voloshinov, who viewed literature as being the effect of current historical situations and social discourses. This functional or causal relation was rejected by the formalists. Marxist attacks on formalism in Russia occurred in two phases. The first was tolerant of formalist endeavours ‘within limits’, labelling their activities as commendable ‘spadework’, with Nicolai Bukharin going so far as to call their work a ‘catalogue’: “A catalogue is only a catalogue. It is a useful thing, all right, but please do not call this inventory genuine science”—quoted by Erlich (1969, 104–5). Leon Trotsky was generally supportive of literary formalists as specialists if their endeavours were restricted to academia, and acknowledged their pioneering gains in literary analysis, but on the question of their claims to a new ‘philosophy’ he was unrelenting, attributing literature to ‘cause’: “Marxism alone can explain why and how a given tendency in art has originated in the given period in history” (Literature and Revolution, 178). In the second phase the Marxist attacks became more visceral as it became apparent that formalism represented a serious threat to socialist realism. Formalism was portrayed by the first Soviet Commissar of Education, A. Lunacarskij, as “a product of the decadent and spiritually sterile ruling class” and as “a relic of the old regime” (Erlich 1969, 107). To Victor Erlich, these pronouncements represented a “dire warning” since the implication was that “formalism was a foreign body in the organism of Soviet society” (ibid.). Pressures on the movement meant it had

5  THE RELATIONAL TURN 

101

largely disbanded by the late 1920s, the final nail in its coffin being the 1934 First Congress of Soviet Writers. Theoretical biologists have endorsed biology as a ‘stand-alone’ science, with properties not reducible to the ‘hard’ sciences. In a similar way to literary formalism, focus is place on the medium itself, rejecting the reconstitution of the biological organism as a means of control in humanist discourse. The organism as an autonomous system is raised by concepts of autopoiesis, auto-catalytic sets and M/R systems. The latter model was developed in one field of theoretical biology, known generally as relational biology.

Relational Biology: Life as a System Relational biologists endorse systemic concepts. Relations both within the organism and with the environment are considered as prior to constituent components (including the nebulous gene). Similarly, as discussed in later chapters, evolutionary developmental biology (evo-devo) has attempted to engage the internal processes of the organism with external, environmental factors. The organism as constituted by complex systems has been the main focus of relational biology. The term, relational biology was first proposed by Nicolas Rashevsky (Fig.  5.2) and taken up by his student Robert Rosen as an appropriate label for the branch of theoretical biology concerned with the organisational features in living cells, as opposed to their constituent components. These are based on complex systems. Rosen links relations to life itself. He states that life is something other than the simple systems that can be explained by mechanics and that ‘relational’ means “throwing away the physics and keeping the organisation” (Rosen 1991, 280). Rosen uses the term ‘complex’ to distinguish life processes from a machine: therefore, a machine like an aeroplane may be complicated, but it is not complex (Cornish-Bowden and Cárdenas 2020). Similarities are found between the autopoietic systems simulated by Francisco Varela and colleagues (Capra 1997, 189) and the M/R systems of relational biology, although Letelier and Mpodozis (2003) put the former within the category of the latter. Both are self-organising and self-­maintaining, not requiring external informational input. Relations in Rashevsky’s sense are more than the relations between components but the effect of these interactions on other interactions. Using set theory, this group of theoretical biologists study these nary

102 

P. MCMAHON

relations, insisting that this adds ‘something else’. Organisms have an abstract quality that cannot be reduced to their components. Furthermore, rather than being a science on the margins of the inorganic world, as Monod (1971) had suggested, studies of life could expand the field of physics itself: “Simple-minded reductionism claims it is a form of vitalism to assert the possibility of learning anything essentially new about physics from a study of organisms” (Rosen 2000, 45). Rosen was inspired by the general systems theory (GST) of Ludwig von Bertalanffy (1901–1972). The phenomenon of protein folding is raised, for example: the activity of proteins is dependent on their 3-D structure, and this cannot be predicted from the primary structure (the sequence of amino acids) or the DNA code. This basic observation is one of many indicators that the phenotype cannot be predicted from its underlying genotype. In ‘Life Itself’, Robert Rosen delineates the processes of life as based on an invariant metabolic relation between production and replacement (or repair, Rosen’s original term). The two systems form a unit (the M/R system—see Fig. 5.3), which does not differ in its basic structure whether present in a bacterium or primate. These relations precede the individual material components which are cycled through the system giving it its Fig. 5.2  The mathematical biologist Nicolas Rashevsky (1899–1972). (Source: Alchetron.com)

5  THE RELATIONAL TURN 

103

Fig. 5.3  A sketch of Rosen’s M/R system. Metabolites S and T are converted to the product ST, catalysed by STU, one of the two catalysts in the system, while an additional metabolite, U, is used for the replacement (or repair) system to regenerate the catalysts (which have a limited lifespan and are subject to decay). The system is materially and thermodynamically open, but according to Rosen, closed to efficient causation. (Note: Catalyst 1 has two functions—multi-­functionality is postulated to be necessary for metabolic closure. Although the catalysts are shown separately here for simplicity, a fuller picture would be provided if they are viewed as forming intermediaries, or complexes which partake in reactions—see Cardenas et al. (2010))

‘structure’. At the heart of this dual structure is ‘closure’, ensured as components of metabolic processes are replaced by another subsystem (the replacement aspect of the whole). This is made possible as some of the material components (reactants or catalysts) overlap between systems—in other words, they display multi-functionality. An Immanent Organisation In relational biology, we find similar themes of self-organisation to those proposed by Immanuel Kant, who contrasted the machine and the living organism. Unlike ideas on life that subscribe to Cartesian mechanism, the

104 

P. MCMAHON

living organism is not a machine. This is because while the parts of an organism depend on other parts to function, similar to the machine in this way, the parts also constitute (produce) the other parts of the organism (Webster and Goodwin 1982). The system is open to materials in the environment, but needs no external direction—it works internally. The machine cannot do this; it is not self-organising, nor self-repairing. Even the computer (our modern reincarnation of the machine metaphor) needs input from an external source (not material input, but informational input) to function. Organisms of course need material input, establishing a paradox of need versus self-­ containment according to Hans Jonas (1966, 76), but they are informationally self-sufficient. There is no DNA floating around that penetrates and informs complex organisms (although a case could be made that naked DNA is taken up and expressed in bacteria, in addition to processes of horizontal gene transfer). Kant differs in this way from his idealist colleagues, who propose that archetypes underly and inform nature. Kant emphasised the immanent nature of the organism. Like evolutionary developmental biology (evo-devo), relational biology is concerned with processes at the phenotypical level. The focus on relations takes biology back to the formalism of anatomists such as Georges Cuvier. For all their limitations, ideas emerging from the Paris Museum of Natural History at the turn of the nineteenth century (Chap. 2) gave no credence to an external agent as determinant of the organism (other than divine design, which at the time was the blanket explanation for all phenomena). Geoffroy Saint-Hilaire was leaning towards the bauplan of the German idealists, but nevertheless focused on the material changes made evident by anatomy, and the homologies they revealed. Cuvier stuck to his observations of anatomy and fossils in drawing his conclusions on the functionality of organisms. A mechanistic empiricism developed in the late nineteenth century after Darwinians suggested a ‘driver’ controls the phenotype, a genetic base that determines the soma, especially after August Weismann’s separation of germ line from somatic cells. To the structuralist biologists, Gerry Webster and Brian Goodwin, the German ‘idea’, on the one hand, and Weismann barrier and ensuing genetic determinism on the other, amount to much the same thing—an external agent (an ‘author’) determines the actual structure of organisms (Chap. 4). Relational biologists present a model close to organicism and self-­ organisation, following a similar theme to Kant, Cuvier and others. However, this model is non-teleological: with no final cause. Relational

5  THE RELATIONAL TURN 

105

biologists are not forgetful of the organism’s relations to its wider environment (and have been an influence on the development of biosemiotics; see Chap. 9) but also claim that the organism, while an open system in the material and thermodynamic sense is self-generating. As well as producing metabolites (reactants and products that, in turn, become new reactants), it also produces its catalysts internally: “Organisms are closed to efficient causation” stated Rosen. This has been called ‘metabolic closure’ by a French group of biochemists who have developed Rosen’s mathematical propositions into more accessible models (Cornish-Bowden and Cárdenas 2008). In other words, the metabolic processes of organism are immanent, with no informational input from an external source (Fig. 5.3). Critiques of Modern Systems Biology Relational biology endorses a holistic model of the functional organism. This branch of theoretical biology challenges current concepts of systems biology (Cornish-Bowden and Cárdenas 2005). Systems biology pays lip service to holism but continues with a reductionist agenda (Cornish-­ Bowden et al. 2004, 2007). Current applications of systems biology have veered away from the general systems theory (GST) of Ludwig von Bertalanffy. They focus on the interactions of components and consequent emergent properties, but are forgetful of the effect of the whole on allocating functions to its parts. Relational biology tries to bring back investigations into the relation of the whole to the function of individual components of a living system. This is supported by discoveries in the field of molecular biology: genes can be edited by post-transcriptional processes, and, therefore, assume more than one function. This is why the number of functional genes is surprisingly low when considering the complexity of metabolic processes, and the number of protein variants needed for metabolic function. The kind of systems proposed in modern biological theory are defined by ‘upward causation’, whereby higher levels of organisation emerge from the interactions at lower levels, resulting finally in a ‘whole’, the intact organism. Systemic properties would include ‘downward causation’, as well (Cornish-Bowden and Cárdenas 2020). Relational biologists begin with the whole, reflected in the title of one of their papers: ‘The parts in terms of the whole’ (Cornish-Bowden et al. 2004). Their model is essentially anti-hierarchical. The evo-devo biologist, West-Eberhard (2003) also raises the bi-directionality observed in phenotypic cellular processes,

106 

P. MCMAHON

but views them as reversible hierarchies. A system relies on a hierarchy but this not always a one-way process: gene function (even of the same gene) may both constitute and be constituted by systemic processes. Thus, biochemists in the relational camp are critical of the ‘reductionist’ systems approach as currently practiced. They state that the system itself is determinant to some extent of its components, and not an emergent property of component interactions. In this sense, relational biology is informed by concepts such as GST. Two key systemic features can be identified: (1) relations in the system precede the actual entities within it, or in other words, components in a system (their position, their function) are determined by a pre-existing relational structure; (2) existing elements assume more than one function, meaning that the systemic character of closure is enabled by multi-functionality. Pertinent to the second point, is that the same elements are used, rather than new ones being introduced (therefore, some genes and their protein products can cross over to operate in other sub-units of metabolism—see Fig. 5.3). Consequently, while living systems are open in a thermodynamic sense, they are metabolically closed, needing no outside direction to function. Relational biology is, therefore, much closer to the ‘operational structuralism’ proposed by Jean Piaget (1971), a transforming and dynamic system based on relations, rather than to a preexisting, static, Gestaltic ‘whole’. It is important to differentiate this from an orthogenetic structuralism, which subscribes to a directionality in evolution. Rather than each gene (or specified DNA sequence) expressing a unique protein and associated function, current research indicates DNA sequences have been modified in evolution—with insertions, deletions, inversions. The gene is like an old building that has undergone various renovations in its history for different functions (office, house, gallery, etc.). Furthermore, genes can be multi-functional. The same region of the genome (or gene) can change its function phylogenetically (in speciation) and can also be multi-functional synchronically. The same transcript (taken from a DNA sequence) can be edited in different ways. Additionally, some proteins have been demonstrated to have more than one function. This lends support to the conclusions of theoretical biologists concerned with relational biology, such as Cornish-Bowden, that metabolic closure requires some components in a living system to be multifunctional. For metabolic closure (a system internal to living cells, with no input other than materials from the external environment), systemic properties must direct the roles or functions of individual components (Letelier et al. 2011). At least some components must be multi-functional to achieve

5  THE RELATIONAL TURN 

107

metabolic closure “[indicating] that a systemic function, closure, imposes multifunctionality on at least some of its components, so this is a genuine systemic property in which a whole system defines properties of its components” (ibid.). Therefore, relational biology introduces systemic principles: rather than systems being constituted by emergent properties, they are also holistic entities—the system is determinant of its components. Multi-functionality in living systems is becoming the norm rather than the exception: A single gene may be both regulatory and structural “such as beta-catenin, which plays a structural role in some contexts and also participates in the transcriptional control of other genes in others” Larsen (2003, 121) states, citing the work of Huber and colleagues. Ellen Larsen continues by pointing out, “most genes do double duty as worker and bureaucratic genes” (ibid., 121). In the field of relational biology multi-­ functionality is particularly pertinent to the notion of metabolic closure. As discussed below, this also addresses the problem of infinite regress. Infinite Regress A problem with catalysis in living cells is that it raises a problem of infinite regress. As we might remember from school biology classes, each set of biochemical reactions in the metabolic system of a cell requires its own set of enzymes. These act as catalysts, raising the rate of reaction. But to synthesise this set of enzymes, another set of enzymes (catalysts) is required, which in turn require other enzymes for their synthesis and so on:

A B CD EF

Enzyme1



Product

Enzyme 2



Enzyme 3



Enzyme1 Enzyme 2

Etc.

This leads to infinite regress: the synthesis of a metabolic product (top reaction) requires an enzyme, which in turn requires another enzyme for its own production, and so on. Rosen’s M/R system and the models developed by Cornish-Bowden and colleagues address this problem. To avoid infinite regress, unless we are to assume that the very last set of enzymes comes from some external source, outside of the cell, the

108 

P. MCMAHON

conclusion reached by Rosen is that the R subsystem (the system that replaces worn out or degraded enzymes) is self-replicating (Letelier and Mpodozis 2003). This could be envisaged if some reactants (metabolites) double up in their roles, acting as catalysts as well (Fig. 5.3). Thus the division between the metabolome and proteome is deconstructed. These biochemists, inspired by ‘relational biology’, reexamine our assumptions about catalysis, which in biology classes are presented somewhat as a black box. Catalysts (proteins, enzymes) are generally presented as external elements that mysteriously appear to ‘help’ each biochemical reaction. Relational biologists suggest that, to the contrary, catalysts themselves enter into reactions creating intermediaries. These molecules are also reactants, even if they are not altered covalently. Relational biologists prefer to view catalysts as simply additional metabolites or reactants, for “there is nothing in the structure of the [metabolic] cycle that justifies a distinction between enzymes and metabolites” (Cornish-Bowden and Cardenas 2007). Catalysts also bind to other metabolites in biochemical reactions, although no covalent change may occur and their molecular structure remains unchanged. This enables biochemical processes to be redefined by relational biologists; reactions now include enzymes as reactants and the complexes they form with other molecules as intermediate products. Additionally, rather than a ratio of one enzyme to one reaction, many enzymes (and therefore the genes that express them) have more than one function. They are part of a complex system that includes multi-functionality. This is evidence of a systemic process at work in living cells that are ‘closed to efficient causation’. The key to organisational invariance is multi-functionality as concluded by Cornish-Bowden and Cardenas (2007): “We have suggested that multifunctionality is an essential part of achieving metabolic closure, or avoiding the infinite regress that is otherwise implicit in the need for every replacement module to have in turn its own replacement module”. Later they state, a doubling up of functions (‘moonlighting’) is “an essential consequence of closure to efficient causation: if every catalyst catalyses just one reaction, and does nothing else, and if every metabolite has just one function, it is impossible to escape from infinite regress. Thus moonlighting is an essential feature of living systems, not just an interesting property” (Cornish-Bowden and Cárdenas 2022, emphasis original). The authors point out that the number of ‘moonlighting’ proteins must be high to fulfil the requirement for closure—to date many proteins

5  THE RELATIONAL TURN 

109

(and genes) have been shown to be multi-functional (Jeffery 2003; Tipton et al. 2003; Gancedo and Flores 2008), lending support to their thesis. As quoted above, Letelier et al. (2011) point out that multi-functionality is indicative of systemic properties at work: it is the system as a whole that allocates components different functions resulting in closure. In other words, the whole system defines the properties of its components. Non-computability Robert Rosen’s work over some decades was largely concerned with putting biology on the map as a discipline in its own right, and as a route to broadening the scope of physics and introducing new principles: I take seriously the possibility that there is no list, no algorithm, no decision procedure, that finds us the organisms in a presumptively larger universe of inorganic systems. This possibility is already a kind of noncomputability assertion, one that asserts that the world of lists and algorithms is too small to deal with the problem, too nongeneric. (Rosen 2000, 3)

While Rosen did not exclude the possibility of artificial life in the future, he believed that present computer technology could not simulate models of life. Rosen was a strong advocate of theoretical biology, which he believed had been largely overlooked since the modern synthesis and the molecular revolution. His studies led him to believe that Erwin Schrodinger’s question ‘What is life?’ and his proposal for a ‘new physics’ had been given short change, with accusations of vitalism. However, the question remains more valid than ever according to Rosen—and cannot not be put to rest by discoveries in molecular biology. Computers in their current stage of development always require an input. Life needs no such input, demonstrating metabolic closure (although as an open system it depends on external materials, therefore is not closed to material causation). This observation has been raised by relational biologists leading to some strident debates (Cardenas et  al. 2010). Life is viewed as non-computable since “organisms are ‘closed to efficient causation’ and that this essential property excludes any possibility of simulable models” (Cardenas et  al. 2010). Simulations to date have been successful, but the catalysts (enzymes) are taken as given. This does not fulfil Rosen’s definition of a simulable model of life, as living cells produce all their enzymes internally. Advocates of artificial life (AL) dispute

110 

P. MCMAHON

the idea that an artificial form of life, similar in all ways to life itself, would be impossible—“the argument about simulability will certainly continue: the work of many groups, including those attempting to develop life in silico, depends on the assumption that computer simulation of living systems is in principle possible, and any claims that it is not possible can expect to meet vehement opposition” (ibid.). To genetic engineers and proponents of synthetic biology, function is an external matter—the function is determined by the element introduced into the system. This contradicts structuralist principles, where function is an outcome of the whole structure, and (in literature) the relations of the dominant (or constructive or foregrounded principle) to other elements in the background. Under the structuralist interpretation, function is aligned with the internal structural arrangement of the whole; inconsistent with ideas that underpin modern biology, where the required (external) function directs the structure. Multi-functionality is one of the basic points in structuralist critiques of genetic engineering. This is particularly relevant since multi-functionality (or moonlighting), as elucidated by relational biology, is intrinsic to the metabolic closure. The problem of pleiotropy has been raised by critics of genetic engineering—introducing new genes inevitably has effects other than the targeted change, as genes may have multiple effects. Genetic manipulation is a heuristic practice, a process of trial and error. Actual physiological effects of the technology are difficult to predict. This might lead to valuable discoveries in the laboratory, but potential problems arise when the environment itself becomes the ‘lab’. An Alternative View Formalism provides an opposing view to current functionalist trends, though it has been actively marginalised—first by socialist realism, but more recently by materialist and reductionist trends under neoliberalism. As Daria Khitrova states: “The rapid transformation of the humanities into a province of the social sciences makes us hope that Formalism is still in our future” (2019, 4). The main formalist attribute, its engagement with the medium of language itself rather than formulating it as a transparent ‘message’, is particularly pertinent to current approaches to biology. Contemporary ‘realist’ approaches (like realism in literary theory) regard the medium (the materials and relations) as mere conduits for functional relations between the external and internal. The word is seen as pointing

5  THE RELATIONAL TURN 

111

to something, rather than an object in its own right. Literary formalists emphasise the way language comes between us and the outer world, for example by its capacity to defamiliarise. In the newly formed Soviet Union, this was not consistent with socialist ideals. Similarly, a biology that is ‘different’ and intractable in its own right is pushed aside by current practises in the biological sciences—the medium is regarded as transparent and unproblematic. The tendency to see biology as an outcome of physical and chemical processes legitimises this approach: understand the chemistry and the physical limitations imposed by membranes and compartmentalisation, and you can understand the system. As Richard Dawkins writes in ‘The Blind Watchmaker’: “My task is to explain elephants, and the world of complex things, in terms of the simple things that physicists either understand or are working on” (1991, 15). The physicist Fritjof Capra points out that the reductionist approach downplays ‘patterns’ (his term for relations): “While it is true that all living organisms are ultimately made of atoms and molecules, they are not ‘nothing but’ atoms and molecules. There is something else to life, something non-material and irreducible—a pattern of organisation” (Capra 1997, 81). Although there are technological challenges, modern biotechnologies proceed on the assumption that information systems in living beings just need to be decodified and, further, that they mirror (map) their conditions of existence. Genetic manipulation falls within the current paradigm: the organism is made of functional components with no inner mode of being, other than the ability to do something, to effect a change. Biology becomes the means to an end.

References Belsey, Catherine. 2002. Critical practice. 2nd ed. New York: Routledge. Brown, Edward J. 1974. The formalist contribution. The Russian Review (Wiley) 33 (3): 243–258. Capra, Fritjof. 1997. The web of life: A new synthesis of mind and matter. London: Flamingo. Cardenas, M.L., J.C. Letier, C. Athel Gutierrez, A. Cornish-Bowden, and J. Soto-­ Andrade. 2010. Closure to efficient causation, computability and artificial life. Journal ofTheoreticalBiology 263: 79–92. Cornish-Bowden, Athel, and M. Cárdenas. 2005. Systems biology may work when we learn to understand the parts in terms of the whole. Biochemical Society Transactions 33: 516–519. https://doi.org/10.1042/BST0330516.

112 

P. MCMAHON

Cornish-Bowden, A., and M.L.  Cardenas. 2007. Organizational invariance in (M,R)-systems. Chemistry & Biodiversity 4 (10): 2396–2406. Cornish-Bowden, Athel, and María Luz Cárdenas. 2008. Self-organization at the origin of life. Journal of Theoretical Biology 252 (3): 411–418. https://doi. org/10.1016/j.jtbi.2007.07.035. ———. 2020. Contrasting theories of life: Historical context, current theories. In search of an ideal theory. Biosystems 188: 104063. https://doi.org/10.1016/j. biosystems.2019.104063. ———. 2022. The essence of life revisited: How theories can shed light on it. Theory in biosciences = Theorie in den Biowissenschaften 141 (2): 105–123. https://doi.org/10.1007/s12064-­021-­00342-­w. Cornish-Bowden, Athel, María Luz Cárdenas, Juan-Carlos Letelier, Jorge Soto-­ Andrade, and Flavio Guíñez Abarzúa. 2004. Understanding the parts in terms of the whole. Biology of the Cell 96 (9): 713–717. https://doi.org/10.1016/j. biolcel.2004.06.006. Cornish-Bowden, A., M.L. Cardenas, J.C. Letelier, and J. Soto-Andrade. 2007. Beyond reductionism: Metabolic circularity as a guiding vision for a real biology of systems. Proteomics 7 (6): 839–845. Dawkins, R. 1991. The blind watchmaker. 2006 ed. London: Penguin. Erlich, Victor. 1969. Russian formalism. History-doctrine. The Hague: Mouton. Fry, Paul H. 2012. Theory of literature, the open Yale course series. New Haven and London: Yale University Press. Gancedo, C., and C.L. Flores. 2008. Moonlighting proteins in yeasts. Microbiology and Molecular Biology Reviews 72 (1):197-210., table of contents. https://doi. org/10.1128/mmbr.00036-­07. Jeffery, Constance J. 2003. Moonlighting proteins: Old proteins learning new tricks. Trends in Genetics 19 (8): 415–417. https://doi.org/10.1016/ S0168-­9525(03)00167-­7. Jonas, Hans. 1966. The phenomenon of life: Toward a philosophical biology. New York: Harper and Row. Khitrova, Daria. 2019. Introduction to permanent evolution. In Permanent evolution: Selected essays on literature, theory and film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 1–24. Boston: Academic Studies Press. Larsen, Ellen. 2003. Genes, cell behavior and the evolution of form. In Origination of organismal form: Beyond the gene in developmental and evolutionary biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 119–132. Massachusetts: MIT Press. Letelier, Juan Carlos, Gonzalo Marı́n, and Jorge Mpodozis. 2003. Autopoietic and (M,R) systems. Journal of Theoretical Biology 222 (2): 261–272. https:// doi.org/10.1016/S0022-­5193(03)00034-­1.

5  THE RELATIONAL TURN 

113

Letelier, Juan-Carlos, María Luz Cárdenas, and Athel Cornish-Bowden. 2011. From L'Homme machine to metabolic closure: Steps towards understanding life. Journal of Theoretical Biology 286: 100–113. https://doi.org/10.1016/ j.jtbi.2011.06.033. Monod, J. 1971. Chance and necessity: An essay on the natural philosophy of modern biology. New York: Knopf. Petersen, Eric L. 2023. A ‘fourth wave’ of vitalism in the mid-20th century? In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T. Wolfe, 173–192. Cham: Springer. Piaget, J. 1971. Structuralism. London: Routledge & Kegan Paul. Rosen, Robert. 1991. Life itself: A comprehensive inquiry into the nature, origin, and fabrication of life. New York: Columbia University Press. Rosen, R. 2000. Essays on life itself. In Complexity in ecological systems series, ed. Timothy F.H.  Allen and David W.  Roberts. New  York: Columbia University Press. Rosengrant, Sandra. 1980. The theoretical criticism of Jurij Tynjanov. Comparative Literature 32 (4): 355–389. Shklovsky. 1992. Art as Technique. In Modern literary theory: A reader, ed. Philip Rice and Patricia Waugh. London; New York, E. Arnold. (Distributed in the United States by Routledge, Chapman, and Hall, New York). Tipton, K.F., M.I. O'Sullivan, G.P. Davey, and J. O’Sullivan. 2003. It can be a complicated life being an enzyme. Biochemical Society Transactions 31 (Pt 3): 711–715. https://doi.org/10.1042/bst0310711. Webster, G., and B.C.  Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-­1750(82)91390-­2. West-Eberhard, Mary Jane. 2003. Developmental plasticity and evolution. Oxford University Press.

CHAPTER 6

Prague Structuralism and the Poetic Function

Structuralists Question Neo-Darwinism and Genetic Manipulation Structuralist ideas in biology became preeminent at the time modernist ideas were challenging the hubris of Victorian ideas of progress and Ferdinand de Saussure presented his Course in Linguistics (Saussure 1959). The ‘structuralist’ biology initiated by Hans Driesch at the end of the nineteenth century arose partly as a reaction to Darwinian and historicist accounts. Darwinian and evolutionary explanations of life forms were challenged. Gerry Webster recognises an opposition in this period between d’Arcy Thompson, Driesch, and William Bateson, on the one hand, and Darwinian theory, on the other: a “system of transformations (a structure) versus a purely ‘empirical’ (temporal or spatial) order” or “law-governed transformation versus accidental or random variation” (Webster 1989). The allusion to transformations reveals that the organism or species was no longer considered the fixed entity of Cuvier’s time. Yet the structural unity proposed by Cuvier was put on the table again, this time as a transforming structure. Contemporary evolutionary theory based on the incremental change part-by-part was criticised by D’Arcy Thompson, for example. Structuralist versions of biology (including relational biology) are more systemic, viewing biological components as functioning according to system requirements.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_6

115

116 

P. MCMAHON

Structuralist biologists have been strong critics of genetic manipulation, which they associate with the adaptationist program and the Central Dogma of one-way flow of information from the gene. Meetings held in Japan in the 1990s by scientists referred to as the ‘Osaka group’ (Kull et al. 2011) were conducted with the aim of identifying risks caused by removing systems of gene expression from their context. Theoretical points were addressed, that is, treating genetic systems as isolated, stand-­ alone units that can be interchanged between organisms and species at will, as well as the effects of the technology, that is, of moving genes into other species and new physiological environments, leading to unexpected pleiotropic effects (Ho 2000). In pleiotropy a single gene may affect more than one phenotypic trait. The heart of the structuralist view is that the whole system is determinant of each part: transformation is not effected by a single ‘part’, such as a gene or protein, but by the whole set of relations. Brian Goodwin was always careful to recognise the importance of the gene, but in the context of the relations that constitute the whole organism: “Genes produce macromolecules, which are essential for morphogenesis. However, an understanding of the sequential action of genes and their products does not provide an explanation of morphogenesis” (Goodwin 1990). Similar themes emerge in biological structuralism to its linguistic counterpart: the organism has structures that develop under certain physical and chemical constraints that only then fit in with the external environment. This reverses the adaptationist stance that external conditions are determinant of structure. Change, it is recognised, has occurred but structures themselves are ahistorical (or have emerged in ‘deep’ history) and last over long periods of time: the backbone is a structure that has changed little, but that does not mean it is a permanent feature of life, just as Saussure’s langue is long-lasting but not written in stone. Change is attributed to the transformation of a set of relations of existing elements, not by introducing new elements or objects. Moreover, proposals in the evolutionary development (evo-devo) camp suggest that phenotypic transformation itself could constitute an evolutionary mechanism (Chap. 8). A transformative notion of structure is endorsed that could be compared to formalist ideas on literary evolution.

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

117

Langue: Saussure’s System Structuralist linguistics arose with modernism in the early 1900s, when Ferdinand de Saussure proposed that language operated as a system, but became especially influential in the 1960s. The term structuralism was introduced in 1929 by Roman Jakobson, who was inspired by Saussure’s linguistic theory: Were we to comprise the leading idea of present-day science in its most various manifestations, we could hardly find a more appropriate designation than structuralism. Any set of phenomena examined by contemporary science is treated not as a mechanical agglomeration but as a structural whole, and the basic task is to reveal the inner, whether static or developmental, laws of this system. What appears to be the focus of scientific preoccupations is no longer the outer stimulus, but the internal premises of the development; now the mechanical conception of processes yields to the question of their functions. (Jakobson 1971b, 711)

Famously, after meeting Claude Levi-Strauss in New  York, Jakobson inspired him to apply structuralist ideas to anthropology. The above quote also indicates Jakobson’s concern with the function of a structuralist system, something that was specifically addressed by Jakobson and the Prague Linguistics Circle (see below). Saussure was interested in examining language synchronically rather than diachronically (the historical, comparative method of the nineteenth century). He viewed the two approaches as separate. Philological (historical) studies looked at change, particularly sound changes in the components of a language. Saussure, agreeing with neo-grammarians that this was an apparently blind (fortuitous) process (perhaps the result of unexpected external events), also observed that this arbitrariness could not be matched with the fact that a language (in any particular period) functioned well, despite these changes (Galan 1979). If there was no system, how could language (beset by arbitrary changes) continue to be functional? To answer this, he attributed functionality to langue (a synchronic set of relations that did not change). Langue was separate from, and took priority over, parole (active speech). In his Course in General Linguistics (1916), Saussure warned that taking the historical, genetic approach of philology, or diachrony, would only falsify a linguist’s “judgement” (Saussure 1959, 81). This, according to F.W.  Galan in his Historic Structures (1985), could be considered the “cardinal insight of

118 

P. MCMAHON

structuralism”, yet it left a system without evolution, a “pitfall” that was addressed by the Prague circle of linguistic theorists. Furthermore, while Saussure proposed that language is a ‘social fact’ he did not develop this theme; the social function of language in communication was taken up later by the Prague school. Language, Saussure maintained, is relational rather than representational of objects in the world. Rather than a word (a sign) being tied to its referent, its meaning is determined by its position in relation to other signs in the system). In structuralist linguistics, entities are determined by their relations or ‘differences without positive terms’. We understand ‘house’ as being not a mansion or a hut, for example, not because we use the referent (an actual house). Meaning is derived from an inner structure, and not ‘out there’ ready to find. Reality (the reality that engages us) is ‘produced’ or influenced by language (langue), and therefore cannot be exactly the same for speakers of different languages. For this reason, human cultures may have different range of colours in their spectrum, not because the object varies (the spectrum of colours is continuous) but because their language divides the spectrum differently—therefore, the  Welsh  word ‘glas’ includes elements of green and blue (Belsey 2002, 34). Different languages may structure concepts slightly differently. The English words referring to water flowing down into a lake or the sea include ‘brook’, ‘stream’ and ‘river’. Stream may be replaced by burn in Scotland or creek in Australia. Those word sounds or marks (signifiers) are connected to slightly different concepts (signifieds) representing different sizes of flowing bodies of water, with brooks being the smallest, and rivers the largest. Other languages separate concepts of bodies of flowing water differently. In French, flowing bodies of water are separated into those that enter the sea (les fleuves) and their tributaries (les rivieres). Therefore, French divides these concepts in a slightly different way to English (Culler 1976, 74). If we only had one English word to describe flowing water, for example, ‘river’, then this would cover all of the above. In that case, the concept ‘river’ would be a broad one, and evoke an image of flowing water of all kinds. In fact, a river (the concept or signified in English) is larger than a stream or a brook and the word ‘river’ signifies the concept of a fairly large body of flowing water. But ‘stream’ (‘a stream of insults emerged from his mouth’, ‘a stream of effluent flowed down the channel’) is also positioned between torrent and trickle. Similarly, it is contrasted with ‘still’ water bodies (lake, pond, reservoir). Every word is positioned according to a series of relations with its ‘neighbours’.

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

119

Language, therefore, partially constructs our world, or ‘speaks the world’. Critically, the meaning of words derives from the contrast or connection with other words in the system. This relation supplies the structure (the rules) of language. A question arises in relation to linguistic structuralism. How does a system of linguistic differences become meaningful? How can it be used to refer to anything in the external world? One answer is that only the concepts that are semantically relevant enter langue so “the identity of meaning is no less a condition of difference … than difference is a condition of identity” (Dews 1987, 28). From a range of possibilities represented by structure, only some are selected based on their semantic relevance. Peter Dews suggests that mediation is required between identity and difference by the speaking subject. In this way Saussurean linguistics also involves an interpretant, just like Augustine’s language theory and Peirce’s triadic sign: “The reciprocal relation between meaning and structure is secured by an interpreting and meaning-­ transforming subject, which seeks to understand itself, to coincide with itself, through language, although this coincidence can never be complete” (ibid., 28). Saussure’s assertion that language is a social fact (a medium) of communication was taken further in Prague—since language had a communicative role, somehow linguistic structure needed to be connected with outer reality (Galan 1979). Here the Prague school took its leave from structuralist purists, and was in turn criticised for this—Louis Hjelmslev believed this was simply a return to functionalism, for example. Nevertheless, structuralist linguistics (especially after revisions in Prague) is all about determining how a linguistic structure can function. Its expressive, aesthetic or autonomous properties cannot be isolated from its communicative, practical (or semantic) roles (ibid.). Saussure placed the synchronic structure and the process of speech (parole) on two different axes, labelled as paradigmatic and syntagmatic. The paradigmatic (vertical) axis is linked to the synchronic, relational structure at the heart of all language, and the syntagmatic (horizontal) axis, is linked to a time-factor as speech (a sentence) unfolds in real time. This model was adapted by Roman Jakobson in his account of the relation of linguistic structure to its contextual applications (see below, Fig. 6.1).

120 

P. MCMAHON

Fig. 6.1  Saussure’s paradigmatic/ syntagmatic axes interpreted by Jakobson as two poles in language: metaphor and metonym. The axis of selection (paradigm, metaphor) is associated with possibilities (a static structure), and the axis of contiguity or combination (syntagm, metonym) with sentences that unfold in actual time

Impact of Structuralism The renewed focus on linguistic structures in twentieth-century philosophy turned subjectivity on its head. Rather than a constitutive and prior subject, the subject after the linguistic turn, in the structuralist movement in particular, was reformulated as a product of language and of discourses of all kinds. Decentering the subject was the key outcome of the renewed focus on language in the philosophical thought that challenged phenomenology and Jean-Paul Sartre’s existentialism in the mid-twentieth century. Following the turn to language as structure, actions could no longer be attributed to individual agents, since combinations, roles, and attitudes were already mapped out by a subject’s place in the symbolic field pertaining to language, culture, political economy etc. The human subject was now regarded as a product, a site produced within language and designated by the linguistic signifier ‘I’. (Goodchild 1996, 113)

Hence, a major change occurred in philosophy following the linguistic turn with a significant weakening of the role of the subject. In general, structuralism provided the theoretical tools needed to deconstruct the subject, which was now postulated to be a product of the ‘system’. But the subject also assumed a ‘humble’ status, and according to Verena Conley, this opened a path for improved environmental relations. As the

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

121

dominance of the subject in philosophical studies was diminished, a way was opened for biocentric relations to assume a place prior to and constitutive of human subjects (Conley 1997). The important message is that these ‘subjects’ depend in the end on the rest of the living world for their existence, something that environmentalists have been saying repeatedly, but to deaf ears. The rejection of the voluntary and unitary subject had repercussions in other areas including Marxist thought. In the 1960s, Saussurean linguistics, representing the key features of structuralism, was taken up by diverse disciplines. The radical move inherent to structuralism shook the humanities—it reversed the assumption of agency, rendering individuals and meaning as products (not agents) of determining structures (Conley 1997). This was expressed in psychoanalysis by Jacques Lacan, in anthropology and kinship structures by Levi-­ Strauss, in modern myths and advertising by Roland Barthes, and by Foucault as discourse and historical formations of knowledge (his famous epistemes). All of these removed conscious agency, replacing it with an unconscious structure (a formation we are not aware of) that produces conscious meaning. Principles of structuralism were applied across the board to various topics. The inauguration of high structuralism was also a symbolic displacement of its predecessor, existentialism. Structuralism provided a major challenge to humanism and Marxist-inspired philosophy, particularly French existentialism, exemplified by exchanges between Jean-­ Paul Sartre and Levi-Strauss. Structuralism influenced literary studies, anthropology, psychoanalysis and the history of science, as well as Marxism. Hence, structuralism expanded its scope beyond language to include social systems, viewing the objects that constitute our social landscape to be constructed objects. They are formed by a play of hidden (unconscious) rules. According to structuralists, these ‘rules’ can be uncovered by scientific analysis. Just as traditional Marxism strives to uncover the ‘rules’ that govern exploitative relations, the way the base (or infrastructure) determines its superstructure, structuralist studies such as the histories investigated by Foucault in his earlier work, tried to uncover the hidden, material regularities and practices that constituted the objects in society. These practices, according to Foucault, are largely discursive. The theme of structuralism is retained in his work, although he has been identified as a poststructuralist. Language and discourse create the concepts, the ‘truths’ we abide by. Similarly, Althusser adopted Marxist scientificity (something he identified with Marx’s later work) to reveal mechanisms of ideology, and how ideological processes fool ordinary citizens, so that they

122 

P. MCMAHON

participate in their own exploitation. A structuralist analysis could be applied to demystify ideology, removing the wool from our eyes. Since concepts are produced by the system of signs (and not by their referents or the outer world) the possibility of meaning being constructed (through sign systems) arises. We seem to be caught up in various constructions of meaning. Language precedes the subject (is external to it) and constitutes it. This notion could also apply to other sign systems: meaning is expressed through the differential relations between signs (so that concepts and ideas are arbitrary, not essential) and not created by autonomous subjects. The common theme in these modes of thought was that hidden rules or common regularities (the structures in question) were responsible for the manifestation, and indeed construction, of diverse phenomena. In anthropology, positioning and regular alignments, including oppositions such as nature:culture, fire:water and kinship relations such as uncle-son and marriage arrangements were postulated as defining relations common to a diversity of social instances in Levi-Strauss’s structuralist anthropology. To Levi-Strauss, oppositions between entities provided the main structural basis for myths and the particular positions and roles of family members in different cultures. Earlier, proto-structuralist thought could be found in Durkheim’s anthropology in which he introduced the idea of oppositions such as sacred:profane as being constitutive of society (Milner 1991). Applying structuralist analyses to popular literature, such as Ian Fleming’s James Bond novels, Umberto Eco proposed that key relations between characters, such as hero:villain, hero:woman under villain’s spell, were repeated whatever the content of the story. It was the structure that defined the narrative (Easthope 1988). Limitations of Structuralism Human individuality, human behaviour and institutional arrangements could be traced to hidden underlying structures. In this sense, the subject was seen as being constructed by the system, not as prior to the system. Rather than observing and recording the external phenomena, the system in question could only be analysed at its deeper level, the level of structure. The orientation of structuralism was quite deterministic and could be likened to the infrastructure (or base) in Marxism, where an underlying structure produces the actual relations, institutions and consciousness in the superstructure. While economic determinism was rejected by structuralism, social formations could nevertheless be analysed by the application

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

123

of theory. Structuralism was regarded as a science, and applied scientific principles that combined empirical studies with theories of signs. It was consistent with the positivist tradition going back to Comte where systems were studied from the outside (Milner 1991). In fact, the positivist aspect of structuralism was one of the major targets for emerging poststructuralist schools of thought. While the movement of high structuralism was relatively short-lived it also challenged function as being prior to structure, proposing underlying structural relations produced the diverse array of phenomena that make up apparent reality (for reality constituted the hidden structures from which the visible was manifested). Rather than function determining the structure, individual structural elements could assume different functions. In the setting of Foucault’s proposed modern episteme, structuralism was truly radical, although it is interesting that three of the big four theorists of structuralism, Barthes, Foucault and Lacan, rejected the structuralist label, leaving only Levi-Strauss as its vocal champion. These thinkers tried to escape the trappings of structuralism, including its determinism and pretensions to scientificity. They moved on to poststructuralist forms of thought, especially after Derrida seriously challenged the premises of structuralism in his seminal paper, ‘Structure, sign and play in the discourse of the human sciences’, presented at John Hopkins University in 1966. Saussure’s separation of language as a synchronic system from diachronic change reflects a major shortcoming of the structuralist model— its static nature and inability to account for change. Revisions by the linguistics circle of Prague in the pre-war years presented language in a new light—as a transformative and transforming structure. Saussure’s dyadic structure of signifier/signified was retained (as opposed to Charles Peirce’s triadic sign), but now the arbitrary changes that affect all languages (and that Saussure believed would undermine his system so should be treated separately) were incorporated into a transforming system that could adjust itself (as a whole) to change.

The Prague Linguistics Circle The Prague circle of linguists met regularly in the pre-war years: they focused on the relation of language structure and meaning, as well as literary theory. Structuralism and formalism, while related, display some differences. The Prague school endorsed the notion of a system, inspired by

124 

P. MCMAHON

Saussure’s linguistics. In literary formalism the function (of a text or device) is already a given. Each component has its role to play and there is little room for ideas of change or evolution. More contextual ideas arose in later Russian formalism, particularly in the work of Tynianov who during a visit to Prague collaborated with Jakobson, introducing systematic concepts presented as eight short theses in a 1928 paper (Tynianov and Jakobson 1972). This paper not only incorporated Tynianov’s ideas on literary evolution, but also Jakobson’s revision of Saussurean linguistics. While Saussure had insisted arbitrary changes could not be considered in langue, which was a complete system with its own rules, Jakobson, recognising that language does change over time, introduced the idea of systemic adjustment. Language can adjust (transform) itself as a whole system to accommodate change: “like any self-regulating system, language is capable of adjusting itself to any ‘blind’ jeopardy” (Galan 1985). Diachrony was therefore incorporated into language, which became a transformative system, rather than a static one. The intention of the individual user (the parole) also received greater emphasis than in Saussure’s account (ibid.). Rather than independent formal devices assuming dominance in a process of defamiliarisation, the idea of foregrounding was introduced within a system of relationships. Elements may be foregrounded (replacing the ‘dominant’ device of the formalists), but they remain in relation with other elements in the system: “This structure is dynamic, containing both the tendencies of convergence and divergence, and is an artistic phenomenon which cannot be taken apart since each of its elements gains value only in a relationship to the whole” Jan Mukarovosky (1932). The Metaphor-Metonym Poles Roman Jakobson, a founding member of the Prague circle after his emigration from Russia, took up Saussure’s counter-positioning of paradigmatic and syntagmatic axes, reconfiguring them into two poles: metaphor and metonym (Fig. 6.1). Krystyna Pomorska (1988, 125) states that these poles were redefined in Jakobson’s work as “fundamental forces acting in language as in all forms of art”. Jakobson connected metonymy to realism, while the construction of metaphor is a literary device associated with the poet (Pomorska 1988, 128).

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

125

The synchronic structure of language is represented on the vertical axis and its utterance in real time on the horizontal axis. Each concept and each sound are clustered with others in a relation of negative difference (Fig. 6.1). As such, each concept assumes its identity as not being another. In the example given above, how can we define the terms ‘river’, ‘stream’ or ‘brook’ in English? The stream assumes a place in the system as a ‘stream’ because it is not a river or brook. Further as a water body, it is not a lake or the sea. Nor is it a torrent or a trickle. This creates an axis of selection: as contiguous words unfold in speech or a message (along the second axis of combination, according to Jakobson) they can be substituted by others from the axis of selection (the code of linguistic signs). Therefore, the words in every sentence have a range of substitutes (in the metaphoric axis). For example, these sentences mean the same thing: 1. The path wound down the slope. 2. The trail meandered down the slope. 3. The track snaked its way down the slope. Grammatically correct sentences are possible by mixing and matching, that is, The track/path/trail wound/meandered/snaked its way down the slope Selection from the paradigmatic axis enables different forms of expression. However, the sentences below would be considered correct grammatically (in their contiguous structure), but out of context: 1. The path walked down the slope. 2. The trail swam down the slope. 3. The track flew down the slope. Therefore, not all descriptive (metaphoric) substitutions are appropriate due to the context of the sentence. In contrast, ‘The car flew along the road’, is accepted as a metaphoric allusion to speed, even though cars do not fly. Jakobson attributed the correct substitutions (those that make sense to the listener) as the result of an understanding of extra-literary context. In the process of selection, the context of the statement is taken into account. In the second set of sentences above, the code is in conflict

126 

P. MCMAHON

with the message since “the syntagmatic, combinative pole is that which anchors language to the pre-linguistic world of events and impressions, while its paradigmatic, selective counterpart is that which effects a more subjective and perhaps bizarre relationship between the mind of the addresser and the code of linguistic signs” (Bradford 1994, 13). The Poetic Function Linda Waugh (1980) recounts in her discussion of Jakobson’s poetic function, that any speech event involves (1) an addresser (encoder, speaker, poet, author, narrator), (2) an addressee (decoder, hearer, listener, reader, interpreter), (3) a code (system, langue), (4) a message (parole, discourse, the text), (5) context (referent), and (6) a contact (“a physical channel and psychological connection between speaker and addressee”). Correlated with these aspects of speech are six functions of which one is predominant “within the verbal message”: (1) emotive (expressive,) (2) conative (appellative), (3) metalingual (metalinguistic, ‘glossing’), (4) poetic (aesthetic), (5) referential (cognitive, denotative, ideational), and (6) phatic, to affirm social contact (Waugh 1980). An example of the conative function provided by Paul Fry (2012) is the verbal message, ‘It is raining’ directed from a mother to a child going out of the door, meaning ‘Put on your coat’, while the same statement (It is raining) can be referential (mere information about the weather), expressive (surprise after a long dry spell) and so on. Of the six functions, the poetic function comprises the focus within the verbal message on the message itself: “[P]oetry is that use of language par excellence in which the dominant function is the orientation toward the message” (Waugh 1980). It is, in fact, “un langage qui met l’accent sur le langage” (ibid., quoting Todorov 1978). This touches on the literariness that the Russian formalists tried to delineate from other textual forms, particularly practical language (the textual form considered ‘non-literary’ by the Russian formalists). The referential comes the closest to practical language in Jakobson’s functional scheme. Its inclusion in the set of speech functions demonstrates that, under the Prague revisions of formalism, practical (standard or ‘transparent’) language was no longer considered to be a separate entity from literariness, but was incorporated into the same system, another entity in the system of relations:

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

127

[P]oetic language has the functions of aesthetic and communicative signs: the literary work, being precisely a set of signs, is both autonomous and communicative. (Galan 1979, italics original)

Literariness now becomes relative—some texts are more ‘poetic’ than others, but none are only poetic or only practical. A multi-functionality is evident as verbal messages incorporate all functions; nevertheless, a hierarchy is established with one being predominant (Waugh 1980). Therefore, the functions are relational so that “in the poetic function, in relation to and as against the five other functions of language, there is a dominance of a focus upon the message” (ibid., emphasis original). Literary Systems as Historical In literature, the two types of text identified by the Russian formalists, practical and poetic language, or literariness (the separation formalists maintained between ‘journalistic-type’ and literary texts) are merged by the Prague linguists. Automatisation (concerned with making language communicative) and deautomatisation (associated with defamiliarisation, aesthetic actualisation, and realising new possibilities) are incorporated into a system of relations. Now “actualisation should be viewed in dynamic, dialectical opposition to automatization” and literature can be studied as part of a whole system, according to “the degree of actualisation of its components” (Galan 1979). In other words, the background of standard language against which the element of actualisation is foregrounded is important—its effect comes from the relation between the two so that “actualization of one element is necessarily accompanied by automatization of other elements” (ibid.). Although more emphasis was placed on the function of the text, its transformation (taking up Yuri Tynianov’s ideas) and contextual influences, the models developed in Prague were still firmly structuralist. Formal devices (i.e., those that defined literariness according to the Russian formalists) could not be considered separately from practical language, but were regarded as participating in a system (Fry 2012). The Prague group, in defining a work of literature, proposed that all structures (devices) should be considered in relation to each other. However, some are foregrounded (replacing the concept of a dominant form in defamiliarisation). Defamiliarisation is the way that literature ‘makes anew’ our perception of reality. It acts so that we see the world with new eyes. However,

128 

P. MCMAHON

this does not mean that other devices are banished, merely that some devices are foregrounded. Now text was a ‘structure’. The notion of foregrounding in the Prague School is a structuralist concept. But the text is arranged according to functional requirements and is strictly relational. These Prague models reflected ideas developed by Tynianov, although he used other terms—his ideas on literary evolution were based on ‘deformation’ as a device assumed dominance (the constructive principle) (Rosengrant 1980). A deformed text is an unmotivated one—this is the basis of change (not a shift resulting in the assumption of a new homeostatic state). The unmotivated system is in a state of tension (Chap. 5). As devices assume ‘equivalence’ (so that the constructive principle no longer ‘deforms’ the structure), the text becomes motivated. A separation between form and content was always implied by the distinction between practical language and literary texts. This was also the case in the New Criticism: New Critics viewed poetic language as dominated by form and practical language by content (Fry 2012, 88). Later developments by Russian formalists, taken up by the Prague circle, dismantled this division, now all was form: For [ the formalists], so-called content is itself a function of poetic language. To put it another way, practical language coexists in any text with poetic language and assumes a function not in relation to reference but in relation to poetic language. (Ibid., 89)

This did not dispel literariness but situated literary aspects of a text with more practical aspects in a structure. All components of a text were seen as in relation with each other—yet some became foregrounded. The revision of both structuralist linguistics and Russian formalist proposals in Prague, involved significant changes in literary theory: (a) context and extra-literary factors were countenanced as important influences on literature, as well as intertextuality; (b) rather than a firm division between literature and practical language, the latter was absorbed into structuralist relations and became another ‘device’; (c) systems could transform themselves, and the synchronic/diachronic division became a theoretical tool rather than an actuality; and (d) the function of devices was not fixed, but could change according to current times.

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

129

Revising the Synchrony/Diachrony Distinction The Prague Linguistics Circle combined formalist elements (inherited from Russia) with Saussurean linguistics, forging a new set of methodologies. Both language and literature change over time and so the system, these linguists maintained, is not a static one. Rather than maintaining Saussure’s distinct diachronic and synchronic aspects, they incorporated the two into a system of transformations. Diachronic change (evolution) of language now had a ‘systematic character’: “[E]very synchronic system has its past and future” state Yuri Tynianov and Roman Jakobson (1972) in a joint 1928 paper, aimed to revise formalist ideas in the light of Saussurean linguistics. A merging of the Russian formalist principle of literary evolution (aptly called ‘Permanent Evolution’ in a 2019 publication on Tynianov’s work) with structuralism put the emphasis on transformative structures. The synchronic/diachronic separation was challenged in the fourth thesis by Tynianov and Jakobson: The opposition between synchrony and diachrony was an opposition between the notion of a system and the notion of evolution. It loses its principal importance insofar as we recognize that each system is given necessarily as an evolution and that on the other hand, evolution inevitably has a systematic character. (Tynianov and Jakobson 1972, 82)

Connection to Extra-Literary Factors In a similar way to Mikhail Bakhtin’s critique of formalism (Oliver 1997, 9), the Prague theorists paid more attention to how literature is affected by context and extra-literary factors. More emphasis was placed on how text interacted with the external world, while prioritising its form, or poetic function. To Roman Jakobson (1896–1982), this function was concerned with the text itself but operated in relation to other functions (including referential, conative, etc.). Linda Waugh states, “Of course, most verbal messages do not fulfil only one function. Rather, they are multifunctional: they usually fulfil a variety of functions, which are integrated on with another in hierarchical fashion with one function being predominant” (Waugh 1980). This presumes a hierarchisation of functions (not a separation of their operation)—they are relational, not absolute categories. It also deconstructs the hard division proposed by the Russian formalists between literary and practical texts.

130 

P. MCMAHON

At first glance the Prague theorists appear to have softened the isolationism (the separation of literariness from other forms of text) of the earlier ideas in Russia, now accounting for context and functional use (how language is used to communicate). But by turning practical language (or content) into another device, literariness was no longer differentiated from content-driven text; it was now given priority over it. Content was simply another element in the system (Fry 2012). Prague structuralism retained its formalist credentials, despite its turn to functional aspects of language. Antony Easthope’s (2002, 11) criticism that Jakobson turned language into ‘coal trucks’ carrying coal to Cardiff (i.e., sender-message-receiver) is forgetful of Jakobson’s poetic function, which is largely acontextual and is central to all other functions. By ‘largely’, not absolutely, acontextual, Jakobson’s structuralism emphasised the relations between all elements in a work of literature—so even the purest instance of the poetic function, the poem, includes contextual elements ascribed to the axis of contiguity or combination: For not only does standard language abound in ambiguity, tension, irony, and every other feature commonly associated with poetry, but poetic language, too, successfully performs the function of standard language, namely, that of referring to extralinguistic reality. (Galan 1979)

The Prague linguists did not place linguistic context in a prior position, as a determinant of language, even with their greater focus on the communicative functions of language. The poetic function was central to Jakobson’s model, the function that engaged with the medium of language itself. Literature draws upon social reality, but is not determined by it or by a direct sign-referent relation: Still, though all features of social reality may play roles in the literary structure, this structure cannot be centrifugal, with single structural elements linking up referentially with particular segments of external reality, but is bound to be centripetal, with all its elements creating a series of internal semantic correlations. Ultimately, if a literary work is to be viewed as a work of art, its aesthetic efficacy must without exception be judged by its structural coherence, not by its ostensible verisimilitude. (Galan 1979)

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

131

Jakobson’s Studies on Aphasia In his interim period in Sweden on the run from the Nazi regime before emigrating to the United States, Jakobson conducted a study on aphasia, identifying two groups of aphasics, correlated with the opposing poles of metaphor and metonym (Fig. 6.1). He found that one group could express sentences that were contiguously consistent, but had difficulty in adding colour or descriptive terms, while the other group produced apparent random, descriptive terms in non-grammatical sentences: [W]e distinguish two basic types of aphasia—depending on whether the major deficiency lies in selection and substitution, with relative stability of combination and contexture; or conversely, in combination and contexture, with relative retention of normal selection and substitution. (Jakobson 1971a)

The two types of aphasia were postulated by Jakobson to be associated with the two poles, metaphor and metonym (Fig. 6.1). One type of aphasic has difficulty in finding descriptive words. For example, the sentence ‘The track snaked its way down the slope’ (see above) might be expressed as: The way went left, then right, to the lower part.

In this case, the aphasic has no difficulty in contiguity (using the correct grammar) but has difficulty in finding descriptive words (from the axis of selection). An example given by Jakobson is that funeral wear is described as ‘what you do [wear] for the dead’—the aphasic cannot not access the word ‘black’ but can allude to the context (Bradford 1994, 16). Similarly, a patient presented with a pencil replies “to write” (Jakobson 1971a). The other type of aphasic appears to readily access descriptive words from the axis of selection (or metaphor), for example, for a track winding down the slope: Winding twisting snakily down

In this case, an incoherent sentence without structure is uttered, while using an excess of descriptive words. This type of aphasic disorder “isolates similarities between words and phrases from their usual relationship with the pre-linguistic context” (Bradford 1994, 17). The disorder is “contexture-­ deficient” and grammar disintegrates into a “word heap”

132 

P. MCMAHON

(Jakobson 1971a): “Word order becomes chaotic; the ties of grammatical coordination and subordination, whether concord or government, are dissolved” (ibid.). The imagist poem with its list of descriptive nouns also draws upon the axis of selection. Observations of young children suggest that before they develop skills in grammar (associated with the contiguity axis) they go through an exploratory stage where they apply all sorts of descriptive words (from the axis of selection). Richard Bradford (1994, 22) quotes Jakobson in a reference to young toddlers: A typical property of children’s speech is an intimate interlacement of two functions—the metalingual and the poetic one—which in adult language are quite separate.

While all the six functions in Jakobson’s scheme are present in relation to each other, his research with aphasics and observations of young children suggest that the axis of selection (the paradigmatic axis) is associated with the poetic function, and in contrast, the axis of combination (the syntagmatic axis) with context and reference. The second axis (contiguity, grammatical sentences) succeeds the first (selection) in a child’s linguistic development, meaning that the poetic function takes priority over the referential. In fact, the second axis of contiguity is an outcome (a selective effect) of the paradigmatic axis. A young toddler’s “initial encounter is always with the paradigmatic axis, more specifically that column of paradigms which relate to his/her own subjective experience rather than to the conventional, normative rules that govern the relation between paradigm and syntagm” (ibid.). The access to an over-abundance of similar words is a stage children need to go through before they can relate their creative thoughts to a contextual situation. Based on this scenario, children do not learn adaptively (by adding more and more to perfectly constructed sentences to make them meaningful) but rather by reducing choices (narrowing down the overproduced choices available in their vocabulary) in forming a contextual sentence. Jakobson’s famous assertion implies the priority allocated to selection (or the paradigmatic axis): The poetic function projects the principle of equivalence from the axis of selection into the axis of combination. (Jakobson 1987, 71)

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

133

Far from being an ‘art for art’s sake’ offshoot of a world defined by science (which Bradford points out was a criticism of formalism and its endorsement of the estranging capacity of literary language), the poetic function, exemplified by the lyric poem, assumes an antecedent primacy in Jakobson’s linguistics, the constituent core of language: “The poet remains faithful to the concept of the linguistic code as the initial point of contact between the self and whatever lies beyond the self” (Bradford 1994, 21). This lends credence to pre-Romantic notion “that all language is originally song-like, subjective and figurative” (ibid., 22). The Flexible Phenotype: The Role of Context in Biology The proposal of phenotypic overproduction has a parallel in Jakobson’s polarity between metaphor and metonym. Organisms relate to their environment efficiently (this is what we can readily observe) but this is the tip of the iceberg. Prior to this, internal choices are made available—it is through the context of particular situations that some are selected while the others are weeded out, or rejected. This entails an overproduction— just as the poetic function relies on an overproduction of connected signs, only some of which are selected through context. It is the relations in the whole organism that provide possibilities for selection. In what is called somatic selection (or developmental selection or epigenetic selection) “large numbers of random, or possibly chaotic variants are produced, and then some variants are selectively preserved or reinforced, while the remainder are unoccupied or eliminated” (West-Eberhard 2003, 37). These are non-genetic processes: “The high degree of flexibility provided by overproduction of variants permits the active, dynamic circumvention of a genetically fixed stable or equilibrium state” (ibid.). These have been called exploratory systems. They apply at all levels from the cellular to behavioural. Charles Darwin observed the process in plants, reported in his ‘The power of movement in plants’ (1880)—he called this circumnutation. A plant such as the Lilium species kept in the dark will display changes in direction of growth until a light source is found (Fig. 6.2). While the reason for this is not completely understood, it improves the opportunity for the plant to detect light. Another example, this time at the cellular level, is the overproduction of microtubules. Microtubules provide the scaffolding structure for mitosis. Microtubules are over abundant in the mitotic cell, being polymerised and depolymerised continuously, but only some are used for functional

134 

P. MCMAHON

Fig. 6.2  Circumnutation, the term given by Charles Darwin to the growth of plants in the dark as they search for a light source. (Isabel Fiorello et al., Creative Commons)

Mean axis of movement Direction

Shape

Amplitude Period

purposes in cell division: “The rapid turnover of microtubules in the mitotic cell generates many different configurations in a short time, most of which are unused” (ibid., 42). The searching behaviours of ants and beetles provide more examples of overproduction cited by West-Eberhard. Army ants set off from a base in all directions in a search for food, before making a single column when a food source is found. Similarly, the male of a species of beetle exposed to a female pheromone moves in apparently random directions in still air, but assumes directionality if the air is move across the space since it can locate the direction of the pheromone source (ibid., 38, 42). These examples of phenotypic plasticity by a mechanism of overproduction suggest they are wasteful in the allocation of resources. But identifying a parallel situation with the poetic function in Jakobson’s linguistics is useful here. An overproduction of descriptive terms (or metaphors) from

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

135

the axis of similarity is a first step in language development of children. Only after this stage is this overproduction contextualised by projection onto the axis of combination (or contiguity). This structuralist interpretation reverses the common-sense view; that is, that grammatical language is learned first, followed by the selecting from a ‘choice’ of metaphors that could be used. Rather the ungrammatical, stage of overproduction comes first. Jakobson observed in both aphasics and children that there is a dissociation between the two axes (the paradigmatic and syntagmatic axes): in children, production of descriptive words takes a prior place to their contextualisation, while one of the types of aphasia identified by Jakobson show a lost capacity to produced grammatical sentences. For this reason, he places the poetic function in a central position in his scheme of the six linguistic functions. Biological phenomena also reveal an overproduction of apparently useless responses and elements, prior to the selection of the most appropriate response. The structural relations that make up the phenotype provide an ‘excess’ of possibilities, an equivalent to the paradigmatic axis of similarity; it is the organism in the context of its environment that selects the most appropriate response. Overproduction allows a more rapid response, evo-devo biologists suggest, than a response relying on gene expression. Research on phenotypic adaptability is probably most developed in neurology, which demonstrates that most synaptic connections are discarded through cell death in a selection process. An overproduction of neural cells and connections is weeded out depending on requirements of the nervous system. But phenotypic adaptability extends to many other responses in both the developing and mature organism. Adaptation, Adaptivity and Phenotypic Plasticity Susanne Langer suggested that language based on internally generated symbols has greater flexibility than models based on external referents (Chandler 2007). If language depended on the naming of every external object, and each affective experience, it would be cumbersome indeed. This also applies to biological adaptation. Rupert Riedl, a critic of the modern synthesis, emphasised the role of evolvability (rather than adaptation)—adaptivity can work both ways as a high degree of variation could be detrimental to a population, but a population that lack variation is less likely to respond successfully to environmental change. In his book, ‘The

136 

P. MCMAHON

normal and the pathological’, Georges Canguilhelm stressed the contrast between adaptation (i.e., being a perfect match for external conditions) and adaptivity, stating that adaptation that is “perfect or complete means the beginning of the end for the species” (quoted by Protevi 2015). In contrast to adaptation, adaptivity presents a range of possibilities. This enables the species to respond to changing conditions. Writing during the heyday of molecular discoveries in the 1960s, Canguilhelm links this range of ‘possibilities’ to Darwinian units of variation. Evo-devo theorists demur, raising the importance of constitutive systems within the phenotype providing it with a flexibility to respond to changing conditions. Adaptivity is linked to phenotypic plasticity (Protevi 2015). Performance in a variable environment is not simply achieved by setting up a genetic effect with a corresponding need or environmental factor. The organism or the species has to be able to respond to changing situations. As opposed to a genetic base (Canguilhelm) or mechanistic responses by phenotypic systems (West-Eberhard), John Protevi turns to the virtual of Gilles Deleuze (and Henri Bergson). The virtual should not be considered as separate to the ‘actual’ but as inherent to it (ibid.). Todd May brings up the example of origami. If the folded paper (the shape) is the ‘actual’, the paper is the ‘virtual’. In the organism, the multiplicity inherent to the virtual, a “pre-individual field of differential relations and singularities” underlies “intensive morphogenetic processes that produce system states” (Protevi 2015). But this system includes both genetic and epigenetic factors, thus avoiding more deterministic interpretations, in which the virtual equates to the germ line, the ‘program’ that is actualised as the soma: “[T]o save Deleuze from tracing empirical individuation back to a transcendental identity qua ‘genetic program’ we must see the biological virtual as the differential Idea of genetic and epigenetic factors” (Protevi 2015). Other structuralist properties could enhance the adaptability of the phenotype. Research in evolutionary-developmental biology (evo-devo) points to the remarkable adaptability of the phenotype. That phenotypes can have alternative states, even in the same life-cycle is clear, and also apparent is that through phenotypic accommodation they may adjust to environmental changes imposed on the phenotypes  (Chap. 8). But less obvious is that the phenotype is capable of producing internal variability (or ‘over-production’) that allows greater flexibility or adaptability and rapid ‘choices’ in response. As suggested above, overproduction in the

6  PRAGUE STRUCTURALISM AND THE POETIC FUNCTION 

137

phenotype has parallels with Jakobson’s poetic function, which he identified as central and constitutive in language development, particularly through his comparative investigation of two forms of aphasia.

References Belsey, Catherine. 2002. Critical practice. 2nd ed. New York: Routledge. Bradford, Richard. 1994. Roman Jakobson: Life, Language, Art. London and New York: Routledge. Chandler, D. 2007. Semiotics: The Basics (2nd ed.), Taylor and Francis 2007 ed. London and New  York: Routledge, Taylor and Francis Group. Original edition, 2002. Conley, Verena Andermatt. 1997. Ecopolitics: The Environment in Poststructuralist Thought. London: Routledge. Culler, Jonathan. 1976. Saussure. London: Fontana. Dews, Peter. 1987. Logics of Disintegration: Post-Structuralist Thought and the Claims of Critical Theory. London and New York: Verso. Easthope, Antony. 1988. British Post-structuralism since 1968. London: Routledge. ———. 2002. Poetry as Discourse, New Accents. Routledge. Fry, Paul H. 2012. Theory of Literature, The Open Yale Course Series. New Haven and London: Yale University Press. Galan, F.W. 1979. Literary System and Systemic Change: The Prague School Theory of Literary History, 1928–48. Proceedings of the Modern Language Association (PMLA) 94 (2): 275–285. ———. 1985. Historic Structures. Vol. Paperback. Austin: University of Texas Press. Goodchild P. 1996. Deleuze and Guattari: An Introduction to the Politics of Desire. USA: SAGE Publications. Goodwin, B. C. 1990. Structuralism in biology. Science Progress (1933-) 47: 227–244. Ho, Mae Wan. 2000. Genetic Engineering: Dream or Nightmare?: Continuum Intl Pub Group; Revised, Updated 2000. Jakobson, Roman. 1971a. Two Aspects of Language and Two Types of Aphasic Disturbances. In Roman Jakobson Seleted Writings Volume II Word and Language. Berlin and New York: De Gruyter Mouton. ———. 1971b. Word and Language. In Roman Jakobson Seleted Writings Volume II Word and Language. Berlin and New York: De Gruyter Mouton. Original edition, 1929. ———. 1987. Linguistics and Poetics. In Language in Literature, ed. Krystyna Pomorska and Stephen Rudy. Cambridge, MA: The Belknap Press of Harvard University Press.

138 

P. MCMAHON

Kull, Kalevi, Claus Emmeche, and Jesper Hoffmeyer. 2011. Why Biosemiotics? An Introduction to Our View on the Biology of Life Itself. In Towards a Semiotic Biology, 1–21. London: Imperial College Press. Milner, Andrew. 1991. Contemporary Cultural Theory: An Introduction. North Sydney: Allen & Unwin. Mukarovosky, Jan. 1932. Poetry and Standard Language (Basnictvi a Spisovny jazyk). In Spisnova Cestina a Jazykova Kultura, ed. Havranek, Bohuslav and Weingart, Milos, 123–256. Translated by FW Galan. Praha: Melantrich. Oliver, Kelly, ed. 1997. The Portable Kristeva. New York: Columbia University Press. Pomorska, Krystyna. 1988. Jakobson Dialogues. Paperback edition (1980, 1983, 1988) ed. Cambridge: MIT Press. Protevi, John. 2015. Canguilhem’s ‘Comparative Physiology’: An Eco-Bio-Social Multiplicity. Symposium: Canadian Journal of Continental Philosophy/Revue Canadienne de Philosophie Continentale 19 (2): 57–71. Rosengrant, Sandra. 1980. The Theoretical Criticism of Jurij Tynjanov. Comparative Literature 32 (4): 355–389. Saussure, Ferdinand de. 1959. Course in General Linguistics. Translated by Wade Baskin. New York: Philosophical Library. Todorov, T. 1978. L’Heritage formaliste. Jakobson, Cahiers Cistre (Laussanne) pp. 47–51. Tynianov, Jurii, and Roman Jakobson. 1972. Problems in the Study of Language and Literature. In The Structuralists: From Marx to Levi-Strauss, ed. Richard T. De George, 80–83. Garden City, NY: Anchor Books. Waugh, Linda R. 1980. The Poetic Function in the Theory of Roman Jakobson. Poetics Today 2 (1a): 57–82. Webster, G. 1989. Structuralism and Darwinism: Concepts for the Study of Form. In Dynamic Structures in Biology, ed. Brian C. Goodwin, Atuhiro Sibatani, and Gerry Webster, 1–15. Edinburgh: Edinburgh University Press. West-Eberhard, Mary Jane. 2003. Developmental Plasticity and Evolution. Oxford University Press.

CHAPTER 7

Immanent Evolution

Literary Evolution as an Internal Process The organised text or living organism faces challenges in accounting for evolutionary change. In fact, in the cases of Cuvier (in biology) and the New Critics (in literary theory), evolutionary concepts are rejected. A fully autonomous literature (the literary poem or text proposed by the New Critics) cannot account for change or evolution. The cohesive organism proposed by Cuvier cannot be changed without compromising functional adaptations to its particular conditions of existence. Formalism has thus sustained criticism for its rejection of historical, diachronic change, that is, of change due to external circumstances and in the case of New Criticism and Cuvier’s embranchements of any change at all, other than new creations that appear suddenly and are isolated (i.e., not linked by common structural features) from other forms, but instead are ‘stand-alone’ independent entities. More sophisticated ideas are needed to account for change. Russian formalism, in its later stages, and the Prague circle of linguistics introduced transformative ideas that could account for a literary evolution. Yuri Tynianov wrote ‘On Literary Evolution’ in 1927 as a response to Leon Trotsky’s ‘Literature and Revolution’ (1926) and “his own quiet protest that evolution is not modification” (Fry 2012). Trotsky’s essay had criticised the formalists for de-linking literary change from revolutionary values. Paul Fry ascribes the ‘quiet’ nature of Tynianov’s protest due to © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_7

139

140 

P. MCMAHON

the increasing oppression of artists and new ideas in the 1920s by the socialist-political machine of Soviet Russia. Expressing a view common to the Russian formalists, Tynianov suggested literature is characterised by a struggle between internal elements, devices that define literariness. He viewed change as being quite sudden and dramatic, as previously dominant forms are replaced by others: forms that were often considered as marginal and already ‘dispensed’ with. Devices or literary forms long-­ forgotten and relegated to the dustbin of literary history can become resurrected, while current forms retreat to the margins. Thus literature is transformed by an internal readjustment, not by an external causal factor (although external factors can directly modify literary production, the political promotion of socialist realism being a case in point). Tynianov distinguishes changes in literary form due to internal transformational processes (an evolution) from those changes he attributes due to particular external events or ideologies, such as the mandate that literature should be put in the service of social or revolutionary values; this he referred to as modification. Yuri Tynianov insisted literature is not the result of periodised movements, such as Romanticism or Realism as “for every current there is a countercurrent, as well as countless competitors to both” (Khitrova 2019, 3–4). Daria Khitrova states in her introduction to ‘Permanent Evolution’ that literary history “is not a history of progress—and that is one of the principal tenets of Formalism” (Khitrova 2019, 13). It is not time-­ dependent (a trundling forward along the axis of time) but a redistribution of values (ibid.). This redistribution occurs quite suddenly, as the periphery, the margins come to occupy the centre. New phenomena do not simply make contributions from the edge but assume a central significance, while older forms retreat to the margins. Therefore, evolution is not linked to historical time but can be jumpy showing “a lack of coincidence, independence and distance” from “real duration” (ibid.). Rather than an evolution of literature, early formalist propositions suggested literature underwent periods of replacement due to the “sudden, complete renovation of formal elements” (Shklovsky 1992). Thus change was seen as sudden, not a consequence of external environmental forces. Viktor Shklovsky’s account was more atomistic than the structuralist version that Yuri Tynianov and Roman Jakobson developed later—Shklovsky considered devices as working independently, while in later ideas they formed parts of a structure, one that was transformed by a rearrangement of its constituent elements. In later developments, Russian formalists also

7  IMMANENT EVOLUTION 

141

became more open to ideas of context and literary evolution, based on the transformation of relations between elements in the system. To Tynianov, while literature was not isolated from external historical forces, it was to an extent also closed: to evaluate “a closed literary series, and to examine evolution within it” always runs up against “neighbouring cultural, social and “real-life” series” (Tynianov 2019, 267). As Khitrova (2019, 4) put it: “Historical phenomena, both literary and otherwise, coexist, though not always at the same time: they work separately or in unison”. But the idea of evolutionary stages, and of a historical determinism, was rejected: “[W]e look for laws of literary evolution within the literary system itself, while not losing sight of literature’s numerous cross points with other orders or series of social reality” (Galan 1979, emphasis original). Therefore, it is not enough to focus only on the verbal art, as in earlier formalist models. But equally, literature should not be seen as a function of historical circumstances: “a sequence either of biographies or of histories of civilization but not a history of verbal art” (ibid., emphasis original). The 1928 joint paper by Tynianov and Jakobson put forward this formalist position in one of their ‘theses’: The evolution of literature cannot be understood so long as the evolutionary problem is overshadowed by questions of episodic, unsystematic literary origins (so called literary influences) and extra-literary origins. (Tynianov and Jakobson 1972, 81)

Modification and Evolution A position common to formalist ideas is the rejection of a causal relation between literary texts and external movements or historico-social factors. Literary formalists from both sides of the Atlantic did not accept that literature is an effect of historical and social conditions; meaning is derived from the text itself (the medium according to Cleanth Brooks) and not by an unfolding historical reality (whether Hegelian, Marxist or neo-liberal). Just as the New Critics rejected extra-literary conditions as a determinant of the ‘text’, the insistence on the autonomy of literature led the Russian formalists to reject external influences as primary forces producing ‘effects’, such as literature. This also suggested that future political events or achievements could not be predicted. Naturally these ideas were not well received by the socialist revolutionaries. Boris Eikhenbaum stated: “Everywhere you look it’s all politics and no scholarship. They will pester

142 

P. MCMAHON

us with ‘causality,’ and rejecting causality will be considered heresy, a deviation, etc.”—quoted by Khitrova (2019, 6). In the Russian formalist movement, just before members halted their inquiries due to Soviet persecution, ideas of evolution of literary forms were put forward by Yuri Tynianov in his ‘On Literary Evolution’ published in 1927 (Tynianov 2019). These ideas differed from the New Criticism, which viewed the literary text as a stand-alone and completed object. Literature, Tynianov proposed, changes by a process of transformation from within—rather than new ideas being introduced pre-existing formal devices are rearranged into new forms. He distinguished literary history from history per se (Fry 2012, 93). Therefore, literature evolves not as result of external inputs, social beliefs, class differences or history, but through an internal transformational process. But literature could also be changed from without; this Tynianov referred to as modification. As Paul Fry points out, Tynianov did not reject modification as a powerful influence in changes to literature, particularly changes resulting from socialist influences, but distinguished such ‘modification’ from the “evolution of literature” (ibid., 92). To the Russian formalists, literature did not evolve incrementally but through radical transformations in the internal structure, or arrangement of forms. As pointed out in Chap. 6, this leads to deformation and tension, a key component of defamiliarisation. But even earlier Russian formalist ideas did not consider texts to be ahistorical. Paul Fry makes the point that the Russian formalists did not ignore history, a charge that is commonly levelled against them. Rather, they rejected the position that literature is a product of “historical-materialist and socialist” influences (Fry 2012, 92). Evolution is first and foremost immanent. It relies on past forms and genres: “According to the notion of immanent evolution, then, a poet is not free to choose any random configuration or structure of elements – but must confront the canon inherited from his or her precursors” (Galan 1985). Literature, the formalists proposed, emerges from a background of prior forms. Defamiliarisation is dependent on change within literature. Literary formalists such as Victor Shklovsky and Boris Eikhenbaum viewed this as a struggle (Fry 2012, 84), as one or other device replaced others as the dominant form. This is because devices that defamiliarise eventually become ineffective, and so are replaced by other devices. But these changes are seen to occur within literature, and not in response to external influences. In his essay ‘Art as Device’, Shklovsky stated: “The purpose of new forms is not to express new content, but to change an old form which has

7  IMMANENT EVOLUTION 

143

lost its aesthetic quality” (quoted by Fry 2012, 93). Ideas on change (literary evolution) and structural transformation were further developed in a collaboration between Yuri Tynianov and Roman Jakobson in their eight theses (see Chap. 6). Biology in its early days as a science also faced resistance to ideas of change. Natural history in the Age of Reason was generally fixist (the position taken by Cuvier) but biological transformism was broached by some thinkers. In opposition to Cuvier, Jean-Baptise Lamarck developed a systematised idea of biological evolution. He did not just raise it as an idea as Erasmus Darwin, Charles Bonnet and others had done, but developed a full-blown evolutionary theory. It was Lamarck’s model that inspired Charles Darwin to consider evolution as a real process needing further elucidation.

Transformism and the Challenge to Fixism The transformism hypothesis was an explanation put forward by the rationalist biologists to explain speciation, predating Darwin’s theory by decades. Darwin himself initially focused on homology and the ideas of Geoffroy (and German Naturphilosophie) based on a unitary body plan (bauplan), only later turning to Cuvier’s studies on anatomy to develop his theory of evolution. The proposal that species transform themselves with time was itself radical to religious orthodoxy and the shift from belief in the immutability of species, God’s creations, to evolution took place against entrenched resistance. The scholastic belief in fixed species can be traced back to Aristotle. Aristotle followed the Platonic idealist form of Greek thought: forms are prior to their manifestation in material nature. Later under Thomism these forms became divine, the blueprints of God. In scholastic thought, the great chain of being was formed by unchanging species created by divinity. Challenging this fixed scheme was to challenge theological belief systems. To Jean-Baptiste Lamarck transformism was characterised by an increase in complexity, from simple infusaria to complex primates (Fig. 7.1). This process was driven vitally. The vital materialism underlying transformism was linked to epigenesis, a concept that explained embryological development by material transformations. In contrast, preformation, supported by orthodox religion, proposed each embryo contained a mini-version of the adult (see Chap. 1). Charles Bonnet and Albrecht von Haller regarded epigenesis as anti-theologian. Ideas of transformation were associated with

144 

P. MCMAHON

Fig. 7.1  Lamarck’s evolutionary theory proposed a vital drive to higher levels of complexity, plus diversification at each stage as individual organisms adapt themselves to changing conditions (Ian Alexander, Creative Commons attribution)

The complexifying force Phyla (Body plans)

Vertebrates Molluscs

The adaptive force Species and genera

Insects

Worms Jellyfish Infusoria

a pernicious materialism. For this reason, early explanations of the development of embryos favoured preformation as opposed to an epigenesis: for the latter theory proposed that matter was mutable and could change its form. Resistance to ideas of transformation from scholastic scholars goes back to conflicts with Renaissance Neoplatonists who proposed matter was mutable and could change from one form to another (Merchant 1980). Alchemy and other forms of ‘magic’ were outcomes of this line of thought. These thinkers also came up against entrenched orthodox religion, some paying with their lives. Resistance to the idea of transformation in pre-Darwinian biology therefore had religious roots. The theology of the time, in common with institutionalised religion generally supported a certain degree of inertness: ideas of transmutation and epigenesis threatened religious preconceptions of nature and the Church, therefore, upheld the notion of the creation of immutable and stable species. It was against the resistance of orthodox religion that ideas of a vital materialism began to take hold. According to John Greene (1981, 42) Lamarck “drew from a thoroughly geological uniformitarianism the inevitable conclusion that organic forms were mutable”. Lamarck proposed that a vital element drove evolution from ‘infusaria’ via molluscs, reptiles, birds to mammals (Russell 1916, 217). The progressive increase in complexity had affinity with German Naturphilosophie and was also taken up by Henri Bergson at the end of the nineteenth century. Naturphilosophie was rooted in idealism, seeking a transcendental anatomy, archetypes and a creative nature. But to Lamarck evolution, while vitally driven, was “a material process of increasing complexity” (Russell 1916, 218). His

7  IMMANENT EVOLUTION 

145

proposal of acquired characteristics accounted for diversification at each ‘level’ of this progressive ‘ladder’ (Fig. 7.1). Proposals of change in the eighteenth century (‘le transformisme’) that predated Darwin’s evolutionary theory were not based on chance nor on contingent environmental conditions, but on incremental changes in structure. Evolution and Orthogenesis Early ideas of evolution endorsed by biologists such as Ernst Haeckel and Jean-Baptiste Lamarck proposed a directionality, or a teleology, progressively leading life forms of increasing complexity up to a final pinnacle, viewed generally as the human being. Steven Jay Gould believed the results of evolution (the phylogenies that have developed) are chance events—they could have taken different turns, and if they had the human species would never have evolved. Contrary to this view, Simon Conway Morris supports a directed evolution with pre-set pathways. In this theory, human-like animals are therefore a likely result of evolution. He is particularly interested in convergence (that is the evolution of similar structures in widely different phylogenies) and gives numerous examples in his ‘Life’s solution: inevitable humans in a lonely universe’ (2003). The modern synthesis of biology, while supporting the apparent trend to complex forms, rejects orthogenesis, or a directional, teleological model of evolution. However, some literary and other figures lament the lack of directionality, wishing for an ordered and progressive model of development. They lament the loss of meaning as Lamarck’s evolutionary ideas were marginalised by the modern synthesis of biology. In ‘The Blind Watchmaker’, Richard Dawkins points out the notion of a non-­teleological evolution, a path with no apparent purpose is unpalatable to some. He points to the widespread emotional attachment to the Lamarckian explanation of adaptation through the inheritance of acquired characteristics (of a directed effort in other words) and a wistful regret expressed at the ascendancy of the Darwinist explanation of a non-directed (apparently meaningless) evolutionary process over Lamarckism. Dawkins cites George Bernard Shaw who lamented “there is a hideous fatalism about it, a ghastly and damnable reduction of beauty and intelligence, of strength and purpose, of honour and aspiration” (Dawkins 1991, 291). Spurning such wishes for directed progress, literary evolution as theorised by the Russian formalists rejects any such historicism or directionality.

146 

P. MCMAHON

In this sense, formalist views come closer to the anti-directionality in evolution expressed by proponents of the modern synthesis. However, rather than ‘blind chance’ there is a systemic basis to the views expressed by Tynianov in his ‘On literary evolution’. Literary evolution is not a question of moving forward, of ‘improving’ but of rearranged structures, with old forms regaining dominance. Tynianov was just as interested in the resurrection of literary forms of the past so that Igor (an old Russian epic) “becomes a modern tale” (Khitrova 2019, 2). As Tynianov stated, “There are roads we have lost, rivers that dried up or changed course” (ibid.). The contemporary and the historical exist side by side: Daria Khitrova quotes Jakobson who pointed out Tynianov’s ‘faith in the coexistence of the present and the past’. She adds: “The present and the past, the contemporary and the historical, do not follow each other like marks etched on a ruler; they coexist, they live side by side” (ibid., 7). Therefore, models (including some interpretations of biological structuralism) that endorses orthogenesis (a directed evolution) need to be distinguished from those that postulate the dynamism of a transformative whole, with no particular teleology or end-point. The latter is closer to the literary evolution envisaged by the Russian formalists and their successors in Prague. Complexity Stephen Jay Gould viewed evolution as an essentially random process, except for a boundary represented by a minimal level of complexity in life forms (viz. the simple prokaryotes, including bacteria). This boundary prevents life forms becoming less complex, so any random ‘path’ will tend over time to veer away from this ‘wall’—the metaphor used is a drunkard staggering in a zig-zag fashion  beside a wall, sometimes recrossing his path, but never breaching the wall. Thus, complexity increases through a passive process by this hypothesis, although at times less complex forms evolve. One example is the loss of eyesight in cave dwelling organisms. Richard Dawkins also points to an evolutionary increase in complexity attributed to a combination of chance events (mutations) and natural selection. One explanation, based on Maxwell’s Demon hypothesis, is that cumulative selection entails the selection of information provided in the environment: “Darwinian selection is a filter, allowing only informative measurements (those increasing the ability for an organism to survive) to be preserved.” Therefore, “information cannot be lost” as “only mutations that reduce the entropy are kept while mutations that increase it are purged” (Adami, Ofria, and Collier 2000).

7  IMMANENT EVOLUTION 

147

In addition, some alternative ideas are based on “the spontaneous emergence of increasing complexity and order” under certain boundary conditions (Capra 1997, 222). Hence according to Stuart Kauffman, biological order is the “result not of natural selection but of the natural order selection was privileged to act upon … Evolution [is] emergent order honored and honed by selection” (quoted in Capra, 221). Self-organisation as observed in many complex systems is key to this natural order; Kauffman has applied Boolean theory to gene networks, to demonstrate possible pathways of cell differentiation. Questions remain. Borje Ekstig asks, if there is a drive towards complexity, why do we still have the chain of being, and Lamarck’s observation of a series of living organisms from simple to complex: “[A]s Lamarck asked already when the very idea of evolution was quite new, if the active power of nature compels life to mount steadily up the chain of being, how can we still see the complete hierarchy today?” and “the oldest organisms of today, although being exposed to environmental contingencies and natural selection for the longest time, nonetheless are amongst the most primitive” (Ekstig 2015). Also, despite claims that life has become more complex, the majority of organisms in the biosphere are the simple prokaryotes (bacteria and archaea). E.O.  Wilson suggested that periods of stasis, found in all taxa, could help to account for this. Long periods of stasis are found in evolutionary lineages or, in other word, periods when species show no change. Formalist thought in literature, with the Russian formalists, and in biology, particularly in some evo-devo hypotheses and alternative interpretations of Darwinism, reject the proposition of directional evolution (or any kind of pre-determined historicism). They also question proposals of a pre-set pathway towards increased complexity. In common with literary formalists, evo-devo biologists point out that complex forms (or sub-­ systems) are ‘recycled’ from the past. Recurrence of atavistic forms occurs in phylogenies (Muller and Newman 2003). This provides evidence that subsystems (or ‘modules’) are not replaced (or banished) but can return in more recent evolutionary lines. These atavistic forms may be simpler than the predecessors—an example is the phylogenetic recurrence in insect orders (referred to as flip-flopping evolution) of three types of ovaries (known as ovarioles): two are more advanced with nurse cells, yet the older type (without nurse cells), first appearing in older evolutionary insect orders, has recurred a number of times in later taxa (West-Eberhard 2003, 367).

148 

P. MCMAHON

Transformation of Structure: Tynianov and Geoffroy Paul Fry equates the idea of evolution in literature to Darwinian evolution (an internal process as mutated genes supersede other genes), suggesting that only major geological or climatic events result in unpredictable modification (Fry 2012, 94). Yet the literary evolution proposed by formalists seems to be closer to the transformism proposed in pre-Darwinian theory, especially by Geoffroy Saint-Hilaire. Darwin’s proposal was that evolution proceeded gradually in response to external changes in the environment, while transformism subscribed either to an internal tendency in life forms towards increased complexity (Lamarck) or to a rearrangement of structural components (Geoffroy). The latter’s idea of saltation is not unlike Tynianov’s rearrangement of internal literary structure. Ideas on saltation emerging in the nineteenth century suggest new types appear suddenly due to environmental changes. Some similarities can be seen here with ideas on literary evolution developed by the Russian formalists: sudden shifts in structural arrangements accounted for literary change. Furthermore, neo-Darwinian theory ascribes evolutionary change to mutations that create modified or new elements. Chance alterations at the genetic level, followed by their selection, enable the development of specialised structures. This differs from structuralist explanations (such as the process of literary change endorsed by the formalists), since these propose that textual forms are inherited from the past, assuming different configurations or arrangements. New forms may also be incorporated, but the main driver of literary evolution is immanent to the system, and not external or causal. Geoffroy believed that evolution occurs due to an internal transformation of structural relations governed by certain laws. Some parallels can be seen with Tynianov’s position, especially in his collaborative work with Roman Jakobson. In structuralist terms, relations between internal entities are altered, allowing the whole to adopt different functions. To Geoffroy, transformation meant that the connections between bone structures were maintained even if those structures assumed new functions. Geoffroy was a child of the rationalist era of eighteenth-century thought, influenced by mechanism and dallying with vital materialist ideas; that is, that spontaneous changes in matter could account for life. However, rather than turning to vitalism (as Lamarck did) he suggested structures are transformed in response to environmental pressures leading to major

7  IMMANENT EVOLUTION 

149

readjustments in structural arrangements (a process of saltation), while maintaining their sequential order. Structures did not evolve to fulfil functions but their own functions could be changed due to structural transformation. This challenges adaptationism, which views structures as an effect (a product) of functional requirements. Geoffroy embraced transformism at a relatively late stage of his career, while Lamarck had been entertaining such ideas for three decades (Appel 1987, 120). Geoffroy was influenced by Lamarck’s ideas of a progressive evolution, but as Michael Ruse states, “for all his favorable talk of Lamarck’s laws, Geoffroy had little interest in the effects of the environment on the adult organism, nor was he much concerned to drive organisms up the Chain of Being through some vague force dependent on needs or whatever” (Ruse 1996, 94). Geoffroy was much more interested in structural transformations—how bones maintained their relative arrangement yet assumed different functions. He was influenced by the German idealists and was in correspondence with Lorenz Oken, who famously suggested that the skull was simply a transformed vertebra (something that Goethe claimed he had thought of before but with little evidence). He was also interested in the environmental effects on embryos, and consequent radically changed forms (monstrosities), of which some could prove to fit new circumstances better than their parents—an early proposal of saltationism or “evolution in jumps” (ibid., 94). Geoffroy’s notion of change has structuralist elements: within a structure, components are rearranged without the introduction of new elements. This was elaborated by Piaget (1971) in his account of structuralism over a century later. The key point is that change is immanent to life forms and not directed externally. Life forms are therefore linked by homologous relations (termed analogous by Geoffroy). This breaks from Cuvier’ four ‘types’ or embranchments, the foundations of our phyla, which are not structurally homologous and have no apparent relation to each other. The dispute between Geoffroy and Cuvier became especially bitter when the former claimed he had identified homologous relations between insects and vertebrates. As mentioned in Chap. 2, the discovery of the regulatory transcription factors, Hox genes, common to both groups lends more credence to Geoffroy’s claim, despite it being intuitively unlikely. Marjorie Grene and David Depew point out that, after the debate, focus was placed on Geoffroy’s unity of composition and homological series, rather than Cuvier’s adapted organism. These ideas on homology

150 

P. MCMAHON

were instrumental to the eventual acceptance of evolution. Cuvier’s static ‘system’ had little to contribute to pre-Darwinian evolutionary ideas, although as Foucault (1970) points out Cuvier’s discoveries became central to Darwin’s theory later. Cuvier’s system comes closer to Saussure’s langue, the synchronous linguistic system of relations that ensures functionality. As discussed earlier, the Prague Linguistics Circle was instrumental in turning Saussure’s system into a dynamic, transforming structure (Chap. 6). Structuralism in biology is modelled on such a system, which raises the possibility of immanent transformation as the basis of change. The irony is that both adaptationism and structuralism appeal to Cuverian biology, but in different ways.

Langue that Incorporates History Structuralist biology has common features with other applications of structuralism, especially Saussure’s linguistic system (Goodwin 1990) and ideas in literary theory. A systemic interpretation of biology might be informed by structuralist linguistics and Saussure’s langue. As suggested already, the revisions in Prague are particularly relevant, as they reconstitute language and literature as dynamic, evolving systems. Similar principles apply as to Saussure’s langue, the basis of structuralist linguistics, which is defined by a structure based on differences without positive terms, irrespective of the particular language or even its stage in history. This set of rules precedes the entities (words) in language. It is definitive of parole or speech. However, differences occur between the structuralisms of the two disciplines—in biology more autonomy is attributed to the parts (entities) of the structure, although the system still has a regulatory function. The linguistics of the Prague school also gives an account of structuralism as dynamic, a system that can absorb change and also transform itself. Intertextuality, introduced by the Russian formalists and an important aspect of Mikhail Bakhtin’s dialogism, is an integral part of the structuralist approach. Also, interchange between literariness and extra-literary factors replaces the firm separation of literariness and non-literature (journalism and texts dependent on history and context) by the formalists. Formal devices (ahistorical to the formalists) were now linked to the context of contemporary beliefs and events: their function can change. A continuity with the past is introduced, with the notion of a literature that evolves, rather than being ‘replaced’ or renovated.

7  IMMANENT EVOLUTION 

151

Saussure’s linguistic system is based on an internal structure of possible relations between signs. However, such a system has to function in a ‘social’ setting; it has to be communicative. The Prague circle of linguists took Saussure’s ideas further, attempting to establish the relation between language structure and function. New ideas on structural/functional properties of language and literature emerged as Russian formalist ideas (some members of the circle were Russian emigres) were combined with structuralist linguistics. The Russian formalists, in earlier days, considered formal devices as separate, independent entities but with the structuralist turn, formal elements, and their relations, assume functions influenced by extra-literary contexts. Literary change was proposed to be based on a radical rearrangement of structure (avoiding a mere continuation of earlier forms, thereby ensuring defamiliarisation). Nevertheless, a connection with past forms was maintained; literary change was regarded as immanent, and not a function of external causes. As discussed in Chap. 8, radical phenotypic changes in biological organisms have also been proposed as an evolutionary mechanism—phenotypic accommodation could be the driver of change, especially in macroevolutionary events, rather than gradual, cumulative changes in a stable population. The organism, Webster and Goodwin (1982) propose, is a structure. This is not a static structure, the langue of Saussure, but a structure that can cope with change—this comes closer to the Prague school model outlined above. Structures maintain themselves and control their elements, but the elements have some autonomy. As F.W. Galan says, in Saussure’s langue the individual in society remains at the margins, looking in. A more dynamic view of langue, however, sees its individual components (its speakers, the practice of parole) as actively maintaining the structure as “for parole, the individual is indispensable, since ‘speaking’ is the prime cause for language evolution” (Galan 1985, 12). Similarly, in biological structuralism, the individual component is not a passive element.

The Organism as a Transformative Structure In linguistics, the separate status of synchrony and diachrony in Saussure’s system was revised by the Prague structuralist school. In language (as revised after Saussure), synchrony and diachrony form part of a dynamic system that is subject to immanent change. Could this apply to biology? Goodwin contends that self-organisation is the key process in the structuralist account of morphogenesis.

152 

P. MCMAHON

To structuralists, organisms “are structural wholes in which the parts are to be understood in terms of their relations to each other and their place in the overall structure”; and therefore, organisms “must be conceived as a reality in their own right” (Webster and Goodwin 1982, 38). In the structuralist account, elements in a system are not independent of their relations. While systems can transform themselves, this does not depend on the introduction of new elements. An example of this is the development of arthropod segments. These form first in an undifferentiated state, and then differentiate (with no new external elements added). In this way, structuralism is opposed to atomism, which proposes a system is an aggregation of parts, called by Jean Piaget “geneticism without structure” (Rieppel 1990). Webster and Goodwin summarise the Piaget (1971) account of the underlying assumptions in structuralist theory: (1) Specific properties are attributed to elements in virtue of the relations they are assigned in the structure: the “development fate is a function of position”, (2) the elements have some autonomy as they too have their own properties, and (3) the whole is transformed, not by changes in its individual elements, but by generating a specific relational structure within certain constraints (such as a law of form). Early models of the organism developed in France, and revived at the end of the century, had structuralist properties, according to Webster and Goodwin, who refer to their “structural and functional unity”, differentiating the organism from the machine. The machine, in contrast, has only “functional unity” (see Chap. 3). Just as in structuralist linguistics, laws of biological form were separated from transformations in time. This structuralist approach appears to have its own (separate) ‘langue’. Goodwin has warned against applying concepts of a pre-existing structure to organisms: A crucial feature of any model of morphogenesis is that it should have the capacity to spontaneously initiate shape formation starting from a spatially uniform state. Otherwise structure is being put into the model initially and this is a kind of historical preformationism that begs the question where this structure came from in the first place. Organisms have to be self-organising, self-generating systems. (Goodwin 1990)

Structuralism in biological thought grants elements with some autonomy, providing a firm material basis for structure. Developmental structuralist models propose the system displays equipotentiality, equifinality

7  IMMANENT EVOLUTION 

153

and self-regulation. For example, a developing chick embryo constitutes a field of potentialities arranged spatially. Various factors have been suggested to determine the behaviour of developing cells, such as chemical gradients and threshold conditions proposed by Alan Turing, for example (Turing 1952), and bioelectric fields (Levin 2012). In the developing embryo, the actualisation of a part (e.g., a wing tip or foot) is restricted by the position of its formative cells in the whole structure, yet experiments have shown that cells outside of this position exhibit the potential to develop into that part if they are disturbed or moved. Also, if cells positioned to develop into a certain part, such as the head of a frog, are transplanted into a salamander embryo, then the cells develop into the head of a frog: they have a functional autonomy, despite being controlled by the relations of the whole system. In addition, other experiments demonstrate the whole structure can recover from perturbation if disturbed. Structural relations in biology can be contrasted with current systems biology—the latter investigates extremely complex interactions, based on empirically isolatable components. Analysis requires high level of computing power due to the massive amounts of information contained in the systematic relations (Cornish-Bowden and Cárdenas 2005). On the contrary, an immanence is suggested in structuralist biology: The biological domain is …. conceivable as a domain which creates itself and within which general and systematic generative processes are at work. (Webster and Goodwin 1982, 26)

Structuralists bring to light structural considerations and physical polarities that are just as important to the survival and propagation of living organisms as genetic inheritance: pre-existing physiological relations, chemical gradients, as well as genes are responsible for structural arrangements and account for organism homology. Adaptive genetic change, it is proposed, occurs in the form of modification, but is not responsible for major structures (such as the vertebrate backbone or the pentadactyl limb). Structural patterns or attractants have been identified in complexity theory: such fields possibly provide the means (and constraints) of organism structure. Structuralist biology emphasises ahistorical processes that affect form.

154 

P. MCMAHON

Morphogenesis Brian Goodwin has proposed that morphogenesis in development may be just as important as natural selection in determining evolutionary form (Goodwin 1994), drawing criticism, particularly from those subscribing to the primacy of genetic processes (Price 1995). However, his work with colleagues such as Lynn Trainor has demonstrated that physical processes underpin morphogenetic fields. Self-organisation of complex systems may occur spontaneously in developing organisms, with no need of genetic direction. While acknowledging that the production of macromolecules depends on genes, Goodwin states “an understanding of the sequential action of genes and their products does not provide an explanation of morphogenesis” (Goodwin 1990). Morphogenesis (the development of shape in organisms), Goodwin proposes, depends on what he calls an “excitable medium”, capable of initiating processes of self-organisation. Morphogenetic fields may be controlled by chemical gradients, bioelectric and other signals that affect mechanical properties (e.g., cell wall flexibility or the cellular cytoskeleton). These are responsible for the organisation of cells (at the level of tissues) and therefore the shape of the organism. They operate both spatially and in kinetic (time-based) processes. Fields are non-local: The quintessential property of a field model is non-locality - the idea that the influences coming to bear on any point in the system are not localized to that point and that an understanding of those forces must include information existing at other, distant regions in the system. (Levin 2012)

Alan Turing proposed that development is controlled by a system of polarised chemical gradients, leading to bifurcation of development pathways. Goodwin expanded this model into morphogenetic fields that can account for three-dimensional shapes. An example is the calcium-­ cytoskeleton-­cell wall system that computer models with pre-set parameters demonstrate can generate cell shape in a self-organising process (Goodwin 1990). Such a model successfully explained the spontaneous formation of the cap of the unicellular marine alga, Acetabularia acetabulum, a domed structure created by turgor pressure on the cell wall, in which flexibility is controlled by calcium concentration and other factors (ibid.). Morphogenetic fields are little understood, but evidence is accumulating that organism form is somehow ‘mapped’, enabling the recovery

7  IMMANENT EVOLUTION 

155

of limbs (e.g., in salamanders) and even the whole organism (from sections of planarian worms). The mechanism is unknown but “perhaps a quantitative model of bioelectrical storage of target morphology will result” (Levin 2012).

Self-Organisation and Homology The focus of early biologists was to elucidate a unitary plan for life (Chap. 2)—the evidence for this was the clear homological relationships between taxonomic groups. It becomes even more striking among embryos of disparate species. Figure  7.2 shows a page from Ernst Haeckel’s ‘Anthropogenie’; it provides a comparison of embryos at different stages of development (although Haeckel has been criticised for exaggerating the similarities). Ernst Haeckel and Karl von Baer came up with different explanations. Haeckel took up earlier evolutionary explanations of

Fig. 7.2  Embryonic development in different species (left to right: fish, salamander, turtle, chicken, pig, dog, cow, rabbit, human) in Ernst Haeckel’s 1877 Anthropogenie, 3rd ed. (APS Museum, Flickr.com). Gill structures are present in all species at an early stage of development

156 

P. MCMAHON

embryological homology (e.g., by the French embryologist, Etienne Serres). In this proposal, there is one universal animal body type and earlier stages of embryonic development reflect the fully developed adult of less complex types. So, for example, the level of development of an adult fish is equivalent to an early embryonic stage of the mammal. Haeckel summarised this as ‘Ontogeny recapitulates phylogeny’. von Baer proposed that the diversity of life is built upon a basic (archetypal) form through various specialisations. The early embryological stage represents this archetypal form, into which new structures are introduced during development. Just as a single foundation of a house can be developed into different architectural expressions in the final product: whether open-plan or with different room arrangements, steep- or flat-roofed, or with French, bay, or Georgian windows, the archetypal embryo, in von Baer’s model, provides the common foundation for the addition of unique, specialised structures. Charles Darwin favoured von Baer’s model over other recapitulation ideas, although his explanation was based on history rather than the transcendental biology of von Baer and others. Despite their differences (von Baer was a vocal critic of Darwinian theory), in both cases they were interested in the introduction of new, specialised structures. This is rejected by structuralists—transformations, they suggest, result from internal alignments of existing elements. Structuralist ideas, therefore, rejected the introduction of new elements into the system. Georges Cuvier straddles the divide here as he believed strongly in the integrity of form, namely, a functional organism based on a structural correlation of parts, a system of relations that could not be changed without reducing functionality. He also maintained strong protestant beliefs, although he never explicitly developed a creationist account. Despite this, the implicit assumption was that God introduced exactly what was needed in forming new species, which Cuvier maintained, occurred in the wake of a catastrophic extinction event. Availability of material for a new organ or structure was not a limitation. Furthermore, there was no room for Lamarck’s transformism (an incremental change with time) in Cuvier’s model. Gerd Muller (2003) proposes that homologues are an instance of organisation of the phenotype, which are maintained for long periods and between different phylogenies. This organisation takes priority over the molecular mechanisms that give rise to them. It follows Richard Owen’s observations of the apparent permanence of some homologies (ibid., 64) and also invokes hierarchical systems that organise their elements. A key

7  IMMANENT EVOLUTION 

157

discovery in recent years is that homologous structures are not necessarily co-extensive with the genetic  pathways that constitute them, indicating that “the position of the homologues in the organizational hierarchy of the phenotype is more important than the pathways of their construction” (ibid., 65). Muller, with Stuart Newman, proposes a three-stage model in the development of homologous structures, beginning with a pre-­ Mendelian stage in which self-organising principles give rise to certain structural arrangements, just as proteins fold into three-dimensional structures spontaneously. In the next ‘Mendelian’ stage, genes are co-opted via genetic accommodation (see Chap. 8) as proposed by Conrad Waddington, although epigenetic processes continue to be important. For example, the famous Hox genes (homologous between widely different taxa) might have been co-opted as “developmental executors” (Willmer 2003, 39). Genes, therefore, become co-opted into new pathways: Many cases are known in which orthologous or paralogous regulatory genes acquire new associations and new developmental roles over the course of evolution. The evolving genome may thus gain control of the epigenetic conditions responsible for the initiation of new building elements. (Muller 2003, 62)

This is followed by the third stage of the process, ‘autonomisation’, when the structure is decoupled from its genetic base. Supporting this are “experimental studies [that] demonstrate that genetic and morphological variation can be poorly correlated” (ibid., 63). There seems to be an interaction between genome and phenotype—developmental genes are co-­ opted and homologues are more permanent than the mechanisms that produce them. A systemic concept is invoked here as “autonomized elements of the morphological phenotype [are] maintained in evolution due to their organizational roles in heritable, genetic, developmental, and structural assemblies” (ibid.). Importance of Stable Traits Structural stability could be important to the evolutionary process. Development constraints are more than restrictive to adaptation—rather, they may be necessary to it. Structures usually work in concert with other structures (something unveiled by Cuvier’s studies in comparative anatomy)—if variation in one structure leads to enhanced fitness, the other

158 

P. MCMAHON

structures would need to be maintained in a stable (invariable) state for this to work. If all of the structures were variable the optimal combinations would be more difficult, something raised by Gunther Wagner.  This is particularly the case for “functionally interdependent characters” (Wagner 1988). For example, molar teeth (useful for grinding) evolved in vertebrates from sharp (canine-like) teeth. Their function depends on attachment to a strong base, provided by the jaw. However, in the selection process, if the jaw bone, attachment structures, and orientation between lower and upper jaws varied, the selection of the flatter surface would not have been advantageous—the constraint of these correlated structures would have been necessary to the evolution of molar teeth according to this hypothesis. Gene Are Ancient Structures Homology also occurs at the molecular level. A structure such as a gene and its DNA sequence is more than its adaptive function; it is an historical structure that in evolution may have been put to more than one use. There is a tendency in the adaptationist school to overlook such historical structures and the constraints they imply in favour of an idealist picture of an organism’s (or gene’s) adaptive function. Many genes are highly conserved between taxa indicating a common evolutionary origin. Orthologs, as they are known (or genes with common ancestry), can produce different transcripts (that code the final protein product) due to alternative splicing (Zambelli et  al. 2010). The expressed proteins, derived from orthologous genes, may have different functions. How can a single (DNA) structure result in the expression of different proteins? One way is to edit the transcript differently through splicing, thus resulting in the expression of different proteins. The first step in gene expression is to make an RNA copy of the target DNA sequence (the ‘gene’). This sequence includes both the useful and useless parts of the DNA, as the DNA was used merely as a template. The next step is to excise the useful parts (known as exons) and splice them together into a new (shorter) strand known as messenger RNA (mRNA). We can picture this by imagining a rope consisting of coloured sections dispersed among black sections—to make the rope fully coloured, we need to cut out the coloured sections then splice them together end-to-end, resulting in a shorter (multi-coloured) rope. In biological terms, the coloured sections are exons and the other sections, introns (Fig.  7.3). Only the exons are used in gene expression of the protein product.

7  IMMANENT EVOLUTION 

159

Fig. 7.3  Alternative splicing of gene sequences: the exons (coloured sections) of the RNA transcript of the DNA sequence (top) are spliced together, but they can be edited in different ways, expressing different proteins. In the version shown at bottom right, the middle section has been omitted (Creative Commons)

Depending on the order that coloured pieces are spliced together, or whether some are left out completely (as shown by the mRNA version on the bottom right of Fig. 7.3), different versions of the coloured rope could be made. This is how mRNA is edited into its final (useful product). The exons are removed from the copied strand and spliced together—this edited, shorter version is used to synthesise the amino acid sequence (the peptide or protein). It might be more accurate to see this edited mRNA as the actual gene, rather than the transcribed DNA sequence. Importantly, the different sections (exons) can be spliced together in alternative ways. Therefore, in a gene expression system a target DNA sequence, through alternative splicing), can result in different proteins. Generally, orthologous genes have similar functions between species but not always. Orthologs with a high percent identity or similarity might produce very different transcripts, therefore proteins. For example, genes coding for chitinases (the proteins or enzymes responsible for breaking down chitins in fungal cell walls) are highly conserved between the mycoparasitic species Trichoderma virens and T. atroviride. In other words, the DNA sequences of the genes in T. virens and T. atroviride are almost identical. Both species parasitise other fungal species. The chitinases in T. atroviride have the function of degrading chitin in the host fungus. This would suggest they have the same function in the sister species, T. virens. But a study on the expression of these genes identified divergent transcripts of the orthologs in the two species and possible pleiotropy (Gruber, Kubicek, and Seidl-Seiboth 2011). This is one example of the common observation of multiple uses of orthologous genetic structures. It points to genes being more than

160 

P. MCMAHON

functional units; many are ancient structures with homologies occurring across taxa but, in a process that is similar to the adaptive radiation observed in homologous anatomies (bird and bat wings, for example), the homologous genes (or orthologs) have assumed divergent functions in evolution. Evidence of homology at the molecular level adds to the anatomical evidence collected in the early nineteenth century of a common ancestry to diverse taxonomic groups. It is also a further indicator that adaptation by organisms to their particular conditions of existence is superimposed on a history of evolved structures. This has implications for the relation between genes (the genetic constitution of an organism) and form. Ellen Larsen (2003, 119) states that “the overwhelming evidence is that that gene families used today are very ancient” and so “we can no longer maintain that new forms evolve merely by evolving new genes”. Goodwin and Brière (1991) propose morphogenetic fields (which can be defined as regulation on cellular processes by tissues) may control development (see above). These fields may explain why “the same genes or cascades” produce different developmental outcomes while different genes may result in the similar morphological outcomes (Larsen 2003, 124). In support of such cross-over functions, many genes have both structural and regulatory functions, indicating a double duty role. An example cited by Ellen Larsen is beta catenin, which acts both as a structural gene and transcription factor.

References Adami, C., C. Ofria, and T.C. Collier. 2000. Evolution of biological complexity. Proceedings of the National Academy of Sciences of the United States of America 97 (9): 4463–4468. https://doi.org/10.1073/pnas.97.9.4463. Appel, Toby. 1987. The Cuvier-Geoffroy debate. Oxford University Press. Capra, Fritjof. 1997. The web of life: A new synthesis of mind and matter. London: Flamingo. Cornish-Bowden, Athel, and M. Cárdenas. 2005. Systems biology may work when we learn to understand the parts in terms of the whole. Biochemical Society Transactions 33: 516–519. https://doi.org/10.1042/BST0330516. Dawkins, R. 1991. The blind watchmaker. 2006 ed. London: Penguin. Ekstig, B. 2015. Complexity, natural selection and the evolution of life and humans. Foundations of Science 20 (2): 175–187. https://doi.org/10.1007/ s10699-­014-­9358-­y. Foucault, Michel. 1970. The order of things: An archaeology of the human sciences. New York: Pantheon Books.

7  IMMANENT EVOLUTION 

161

Fry, Paul H. 2012. Theory of literature, the open Yale course series. New Haven and London: Yale University Press. Galan, F.W. 1979. Literary system and systemic change: The Prague school theory of literary history, 1928–48. Proceedings of the Modern Language Association (PMLA) 94 (2): 275–285. ———. 1985. Historic structures. Vol. Paperback. Austin: University of Texas Press. Goodwin, B.C. 1990. Structuralism in biology. Science Progress (1933–) 47: 227–244. Goodwin, Brian C. 1994. How the leopard changed its spots: The evolution of complexity. London: Charles Scribner’s Sons. Goodwin, Brian C., and Christian Brière. 1991. Generic dynamics of morphogenesis. Greene, John C. 1981. Science, ideology and world view. University of California Press. Gruber, S., C.P. Kubicek, and V. Seidl-Seiboth. 2011. Differential regulation of orthologous chitinase genes in mycoparasitic Trichoderma species. Applied and Environmental Microbiology 77 (20): 7217–7226. https://doi.org/10.1128/ aem.06027-­11. Khitrova, Daria. 2019. Introduction to permanent evolution. In Permanent evolution: Selected essays on literature, theory and film - Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 1–24. Boston: Academic Studies Press. Larsen, Ellen. 2003. Genes, cell behavior and the evolution of form. In Origination of organismal form: Beyond the gene in developmental and evolutionary biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 119–132. Massachusetts: MIT press. Levin, M. 2012. Morphogenetic fields in embryogenesis, regeneration, and cancer: Non-local control of complex patterning. Biosystems 109 (3): 243–261. https://doi.org/10.1016/j.biosystems.2012.04.005. Merchant, C. 1980. The death of nature: Women, ecology and the scientific revolution. USA: Harper and Row. Muller, Gerd. 2003. Homology: The evolution of morphological organisation. In Origination of organismal form: Beyond the gene in developmental and ­evolutionary biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 51–70. Massachusetts: MIT Press. Muller, Gerd, and Stuart Newman, eds. 2003. Origination of organismal form: Beyond the gene in developmental and evolutionary biology. In The Vienna series in theoretical biology, ed. Gerd Muller, Gunter Wagner and Werner Callebaut. Massachusetts: MIT press. Piaget, J. 1971. Structuralism. London: Routledge & Kegan Paul. Price, C. 1995. Structurally unsound: A review of how the leopard changed its spots: The evolution of complexity, by Brian Goodwin. Evolution 49 (6): 1298–1302.

162 

P. MCMAHON

Rieppel, Olivier. 1990. Structuralism, functionalism, and the four Aristotelian causes. Journal of the History of Biology 23 (2): 291–320. Ruse, M. 1996. Monad to man: The concept of progress in evolutionary biology. Cambridge, MA: Harvard University Press. Russell, Edward Stuart. 1916. Form and function. Shklovsky. 1992. Art as Technique. In Modern literary theory: A reader, ed. Philip Rice and Patricia Waugh. London; New York: E. Arnold. (Distributed in the United States by Routledge, Chapman, and Hall, New York). Turing, Alan M. 1952. The chemical basis of morphogenesis. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 237 (641): 37–72. Tynianov, Jurii. 2019. On literary evolution. In Permanent evolution: Selected essays on literature, theory and film  - Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 267–282. Boston: Academic Studies Press. Tynianov, Jurii, and Roman Jakobson. 1972. Problems in the study of language and literature. In The structuralists: From Marx to Levi-Strauss, ed. Richard T. De George, 80–83. Garden City, N.Y: Anchor Books. Wagner, G. P. 1988. The influence of variation and of developmental constraints on the rate of multivariate phenotypic evolution. Journal of Evolutionary Biology 1 (1): 45–66. Webster, G., and B.C.  Goodwin. 1982. The origin of species: A structuralist approach. Journal of Social and Biological Structures 5 (1): 15–47. https://doi. org/10.1016/S0140-­1750(82)91390-­2. West-Eberhard, Mary Jane. 2003. Developmental plasticity and evolution. Oxford University Press. Willmer, Pat. 2003. Convergence and homoplasy. In Origination of organismal form: Beyond the gene in developmental and evolutionary biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 33–49. Massachusetts: MIT press. Zambelli, Federico, Giulio Pavesi, Carmela Gissi, David S. Horner, and Graziano Pesole. 2010. Assessment of orthologous splicing isoforms in human and mouse orthologous genes. BMC Genomics 11 (1): 534. https://doi.org/ 10.1186/1471-­2164-­11-­534.

CHAPTER 8

Phenotype then Gene

Evolution by Combining Traits As discussed in the previous chapter, both literary and biological evolution have been proposed to proceed by the combination of pre-existing forms or traits. In biology, the notion of such a phenotypic process challenges theories based on the isolation of populations (and their genes) as the main driver of evolutionary change. Dynamic principles emerge in evo-­ devo discussions: combinatorial evolution (where pre-existing modules are recombined), the recurrence of atavistic structures in later phylogenies, multi-functionalism (where the same gene functions in more than one sub-system) and the idea of phenotypic plasticity. Phenotypic plasticity can result in morphic types developing under different environmental conditions, or different nutritional regimes. Therefore, a lake fish, the Arctic charr (Salvelinus alpinus), has three distinct forms that feed on either plankton, macro-invertebrates, or fish (Crispo 2007). Behaviour can also be plastic. Generally, explanations for change invoke the whole organism, or at least systems within it and their constitutive relations, although mutations and new elements are not excluded as possibly important factors. A more dynamic picture emerges, compared to one that depends on constraints, which comes under scrutiny: “The traditional emphasis on stability, homeostasis, equilibrium, canalization and resistance to change has its modern counterpart in the concept of ‘developmental constraints’” (West-Eberhard 2003, 8). © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_8

163

164 

P. MCMAHON

Population biology, while providing good explanations for adaptive change in closely related populations and species, has less to say about the evolution of different body plans. It is hindered by what Mary Jane West-­ Eberhard calls “the stability obsession” (West-Eberhard 2003, 8). Due to this insistence on stability, polymorphism is disputed as an evolutionary mechanism. Yet intraspecific polymorphism is the norm in biology: juveniles, Queen and worker bees, alternate hosts in parasitism, and metamorphosis, being examples. Polymorphism is particularly striking in many life cycles, showing heterochrony or variant structures. Many adult sessile life forms (sponges, for example) have actively swimming larvae as their juvenile stages. Therefore, heterochrony, according to Stephen Jay Gould, could be a powerful source of evolutionary change (Gould 1977). Modularity is particularly important as it maintains that internally consistent traits are in place that can be switch-controlled and turned on and off (see below). This enables novelty to arise based on a pre-existing phenotype (West-Eberhard 2003, 200). Internal transformation may produce novelty: “Phenotypic variants are to a degree products of history, but they are not bound by the past history recorded in the genome, for the numbers of potential new developmental rearrangements are enormous, and they can be either mutationally or environmentally induced” (ibid., 374). Under the modern synthesis, the idea of cohesiveness presents a problem for speciation. A population becomes “a single, stable or slowly oscillating mode” (ibid., 8). Therefore, isolated populations prevent gene mixing with other populations leading to a new central point—something that West-Eberhard calls “unimodal adaptation”. It is reproductive isolation that disrupts the stable mode of populations: “the disruption of cohesiveness by speciation is still considered by many to be a requirement of evolutionary change” (ibid., 10). Based on the Hardy-Weinberg equilibrium, unimodal adaptation in population biology has promoted a homeostatic view that populations resist change, and “[t]hus polymorphism came to be seen as unstable, and (except under special circumstances such as heterozygote advantage or frequency dependence) selection is expected to produce a single adaptive norm” (ibid., 7). In contrast to the population thesis, “[c]ombinatorial evolution, or evolution by reorganization of preexisting traits, is possible because of the modular quality of phenotypic organization and because the switches that control development are subject to influence by numerous genetic and environmental cues” (ibid., 200). This is supported by various evo-devo studies and observations on

8  PHENOTYPE THEN GENE 

165

phenotypic plasticity, alternative phenotypes (including polyphenism), recurrence and modularity. The gene is downgraded as a major player in the evolution of novel traits, but is still the most important instrument of heredity (the mechanism by which selected traits are passed on to the next generation). This entails a linkage being formed between phenotypic and genetic components: Plasticity is what makes possible the appearance of an environmentally induced novel phenotype, and a process of selection on the expression of such phenotype in a new environment may end up ‘fixing’ (genetically assimilating) it by altering the shape of the reaction norm. (Nanjundiah 2003)

These ideas are particularly relevant to macroevolution, or the emergence of novel types. At the population level the modern synthesis provides the best explanation for variation and the distribution of genes through a population. At this level, genetic recombination contributes to variation, as well as ensuring genes are spread through the population. However, West-Eberhard regards genetic recombination as “ephemeral” since genes for complex traits often occur on different chromosomes and “are continuously crossed over and mixed”. This is quite a random process, and one that is not reproducible so “is difficult to envisage as the basis of a novel trait” (West-Eberhard 2003, 200). As discussed further below, genetic accommodation may result in a phenotypic trait being fixed either by genetic assimilation or by retaining environmental sensitivity (leading to polyphenism, for example). The notion that developmental variation governs evolution is the reverse of Ernst Haeckel’s biogenetic law of recapitulation—rather than ontogeny reflecting phylogeny it is the other way around: Haeckel’s Biogenetic Law [was] summarized by the famous phrase “ontogeny recapitulates phylogeny,” as if phylogeny were the causal agent and ontogeny the result. Instead, there is a sense in which phylogeny reflects ontogeny: development gives rise to the variation that underlies phylogenetic diversification. (West-Eberhard 2003, 373)

166 

P. MCMAHON

Going Further Than Constraints and Self-Assembly Neither constraints nor self-organisation is enough to explain the evolution of diverse forms according to evo-devo biologists, such as Mary Jane West-Eberhard. Stuart Kauffman has proposed a model of development based on attractors or self-organising entities characteristic of physical systems. To Kauffman, “development [is] part of a more general class of phenomena governed by systems with their own “inherent” spontaneous order or self-organization” (West-Eberhard 2003, 128). In addition, some structuralist ideas turn to physical processes, such as reaction-­ diffusion in the phenotype to explain development (Chap. 7). Spontaneous self-organisation, pan-environmentalism and physico-structuralist proposals side-step “the problem of relating genomics and environmental influence in development” (ibid., 99). In evo-devo, both sides (environment and gene) are viewed as critical which, in its way, deconstructs the nature-­ nurture debate: “If genetic and environmental influences are equivalent and inter-changeable, they are not properly seen as opposed or even as complementary factors” (ibid.).

Sudden Evolutionary Shifts Proposed by Biologists: Mutationism Some biologists at the turn of the twentieth century did not regard Darwin’s theory of gradual, cumulative variation as a sufficient explanation for the variety of formal structures or body plans found among biota. Ideas of mutationism (meaning major sudden change) were raised, reminiscent of Geoffroy Saint-Hilaire’s earlier saltation thesis that evolution is a result of environmental pressure on embryo development. However, those endorsing mutationism did not reject a Mendelian (genetic) explanation for sudden change—they embraced it. Biologists including Hugo de Vries, William Bateson and Thomas Hunt Morgan took Gregor Mendel’s discoveries to support ideas of mutationism—for Mendel’s research had demonstrated particulate heredity. Hugo De Vries compared structures in wild species of primroses, suggesting new species had emerged suddenly. Such ideas challenged Darwin’s idea that evolution is gradual. Therefore, in the early twentieth century, during the ‘eclipse of Darwinism’, mutationism questioned Darwinian theory and was even opposed to it (Dawkins 1991).

8  PHENOTYPE THEN GENE 

167

However, the modern synthesis provided an explanation that accommodates mutations (as small genetic changes) with gradualism. It was Ronald A.  Fisher who showed that heredity is indeed particulate, but based on very small ‘particles’ (genes) often with additive effects. This is why we can dispense with early objections to Darwinian theory, that of ‘swamping’. To simplify the idea of swamping (or blending), this pre-­ Mendelian idea suggested if a black horse was crossed with a white horse producing grey progeny, black or white horses could not be produced from grey horses in the next generation. The metaphor used is if two paints (black and white) are mixed you cannot recover the two original colours from the now grey mixture. But genes are not paint—they behave in a particulate manner, and can be recombined providing a rich source of variation. This is the basis of Mendelian genetics and also quantitative breeding methods (adopted to develop new crop or livestock varieties). The basic principle used by agricultural breeders is that complex characters are determined by many small genetic (particulate) effects. It is true as Richard Dawkins points out that random major mutations are not likely to be conducive to survival. However, new ideas on phenotypic plasticity and accommodation look more closely at radical change as possibly being a creative evolutionary force. First, not all changes need to be mutational via the gene—environmental pressures can lead to alternative phenotypic states. Second, a new phenotype (a monster) might be accommodated by changes in the form of the system—this may be possible due to phenotypic plasticity. Some believe this provides a working hypothesis on how macroevolution (and speciation) is possible. Studies on phylogenetic changes in taxonomic trees support the idea that speciation occurs rapidly, perhaps even the result of a single, rare event. In one meta-analytical study, determinations of the branch length (signifying length of time) between 101 phylogenies determined that 78% of the speciation events were accounted for by single (but rare) changes; the hypothesis is that major changes or mutations lead to reproductive isolation and therefore speciation (Venditti et al. 2009). By turning to evolutionary models of sudden change envisaged in literary evolution and speciation in biology, alternatives to Darwinist gradualism can be considered. In addition, both these literary and biological evolutionary models reject the notion that evolution is directional. Those who criticise the notion of random change seem to want to replace it with something. What? One is an inherent drive towards increasing complexity, ending up with mammals and humans. This, of course, was Lamarck’s

168 

P. MCMAHON

central evolutionary idea (while the inheritance of acquired characters provided an explanation for diversity within phylogenetic groups). But the literary formalists, in comparison, rejected any notion of cause or directionality. This accounts for their disfavour with the Stalinist regime of the time. Similarly, it explains the marginalisation of formalist ideas in current biology  by neo-liberal win-win ideals  pursuing  a humanist ‘cause’: help the world by modifying life forms and make profits at the same time. Tynianov’s scheme is much more akin to a dive into the past, a reclamation of lost ideas and genres, a ‘recurrence’ in fact, just as phylogenetics has demonstrated in some taxonomic lineages, whereby traits, rather than evolving into new forms, reappear in later phylogenies. The transformations underlying literary evolutionary as proposed by Tynianov, and similarly in alternative biological models, are not completely random, but nor are they directional or predictable. They are not completely random as they are accessing ancestral or archival material—that had at one time its own functionality. But where and when they emerge depends on both internal and external chance events. Darwin himself did not dismiss saltation as a viable alternative explanation for evolutionary change, but he favoured the gradualist thesis. The modern synthesis has reinforced and reaffirmed gradualism, dismissing other ideas such as the ‘hopeful monsters’ suggested by Richard Goldschmidt. Now with the advent of evo-devo, ideas such as saltation and phenotypic adaptability are making a come-back. Some of these ideas are discussed below, relying heavily on Mary Jane West-Eberhard’s seminal publication, ‘Developmental plasticity and evolution’ (2003). The modern synthesis, largely turning to population biology and its component genes, has dominated evolutionary debates, but evo-devo biologists have turned their attention to the phenotype by looking “beyond mutation to seek the origins of variation in the developmental plasticity of organisms” (West-Eberhard 2005a).

Punctuated Equilibria The fossil record (e.g., in the famous Burgess shales) shows long periods of stasis maintaining a certain ‘bauplan’, but over relatively short periods rapid transformation results in the production of diverse forms. Niles Eldredge first proposed that rather than phyletic gradualism, evolutionary diversification of species occurs in spurts followed by periods of stasis (Eldredge and Gould  1972). Explanations of how original (stable)

8  PHENOTYPE THEN GENE 

169

structures emerged are lacking in evidence. However, Stephen Jay Gould referring to the Cambrian explosion points to fossil evidence in the Burgess shales of a rapid divergence in primary forms—many later became extinct. The term punctuated equilibria was coined to describe this process, counterposing it to gradualism. Although Gould concurred allopatric speciation based on Darwinian natural selection could be the mechanism, he pointed to the inconsistencies of this theory in explaining macroevolution. He broached other ideas, such as arrested heterochrony (alternative phenotypes and immature forms in the life cycles of many animals). Gould views natural selection as a fine-tuning of existing organs, and Brian Goodwin in agreement thinks selection is important at the population level, but offers a poor explanation for macroevolutionary events, such as speciation. Multi-level selection is raised as an alternative to selection of favourable genes—thus units of selection might occur at the population, clade and species level.

Modular Systems The differences between genetically identical individuals (e.g., twins or cloned plants) has been put down to developmental ‘noise’, but evolutionary developmental biologists such as Vidyanand Nanjundiah demur. Often it is an instance of phenotypic plasticity and this plasticity gives the organism a capacity to adapt (Nanjundiah 2003, 244). A point Nanjundiah raises is that since environments are unpredictable, there may be “no single optimal phenotype” (ibid., 253)—plasticity becomes particularly important in this regard. With the rise of evo-devo there is an increasing focus on post-­ transcriptional processes. Individuals differ due to both genetic and epigenetic factors. Variation in gene expression accounts for cellular specialisation and different tissues, despite the same genome being found in each cell. Physical and chemical factors (such as gradients defining fields) and alterations to DNA (structural alterations to chromosomes and methylation, for instance) contribute to cellular processes. Furthermore, pathways of gene expression may be modular: As we now know, there are “developmental networks” that are used differently in a wide range of tissues even within the same organism, which suggests such networks are modular, and that modules can be snapped in or out and can be redeployed when not needed for a particular task, with ­widespread

170 

P. MCMAHON

redundancy of genetic (Willmer 2003, 38)

pathways

in

developmental

processes.

Modularity is now widely accepted among evo-devo biologists (Hall 2003). Stable sub-systems are more important to evolution than stable components (the molecules themselves). Ellen Larsen (2003, 126) recounts a parable of two watchmakers. Two watchmakers are given the task of assembling a watch from 1000 initial components, but as they do this they are interrupted a number of times, for example, by a phone call, resulting in their work falling apart. One watchmaker tries to assemble the watch by combining components individually, while the other first constructs smaller ‘subsystems’ consisting of 100 components (we could call these modules), which are then assembled at the end. Due to the interruptions the first watchmaker cannot complete the task (the watch falls apart at each interruption) while the second watchmaker easily finishes the task due to the completed 100-component subsystems, each of which are stable. If the components of the watch are metaphors for genes, and natural selection was to operate in this scenario, the modular form would allow evolution to proceed more rapidly than evolution based on individual genes (ibid., 126). The gene cascade could allow modularity by modification of its initiation at the receptor. Environmental cues could affect key regulators, causing a switch in developmental pathways. Embryological studies reveal discrete regions, groups of cells, that act as morphogenetic fields, the precursors to developing structures—for example, the tetrapod limb bud (Raff 1996). The field is integrated so that individual (dissociated) cells cannot replace the function of the cell group. These fields are instances of modularity, characterised by “discrete genetic specification, hierarchical organization, interactions with other modules, a particular physical location within a developing organism, and the ability to undergo transformations on both developmental and evolutionary time scales” (Bolker 2000). Modularity might be the explanation for independent populations of a species of butterfly assuming identical wing patterns. Mary-Jane West-­ Eberhard suggests that recurrent wing patterns in the variants of Heloconius butterflies, located as isolated populations in eastern and western South America, are evidence of modularity, rather than convergence: “This enabled the somewhat isolated populations of the same species to undergo the same kind of change independently in different descendant sublineages” (2003, 357). She calls this “recurrence homoplasy”. Many

8  PHENOTYPE THEN GENE 

171

so-called instances of convergent evolution (when the same trait evolves independently due to similar natural selection pressures) might be, in the end, attributed to the recurrence of a modular (cryptic) system. Modular sub-systems, if expressed, could be subjected to natural selection. Therefore, combinations of existing (cryptic) traits could also lead to changed wing patterns in butterfly species. Natural selection could then enable the new phenotype to spread—followed by genetic accommodation (see below). This means that natural selection might operate in both scenarios: step-by-step adaptation based on cumulative mutations, or changed phenotypic states, based on pre-existing systems. In other words, Darwinian natural selection, particularly at the population level is not rejected as an important adaptive mechanism, but the re-emergence of cryptic traits in phylogenies is put forward as an alternative mechanism. Nevertheless, an issue remains on where and when the (modular) cryptic traits (in the case of phenotypic adaptability) evolved in the first place. If evolution can be explained by descent with modification, then major conserved forms and complex structures remain difficult to explain. The ‘descent’ aspect, or homology (the remarkable similarity between different species), requires that at some point these structures evolved. Darwin himself warned against applying natural selection to explain all biological form. Subscribers to the modern synthesis claim to have come up with robust answers, based on the frequency of adaptive genes in populations and their isolation (e.g., by geographic barriers) leading to formation of new species, but these ideas while useful in population biology are not universally accepted as an explanation of speciation. This is where some biologists part from the adaptationist paradigm. Whilst neo-Darwinism provides an account of adaptive biological structures that have arisen under particular conditions and certain times when necessitated by environmental condition, leaving a legacy for following generations (the ‘descent’), greater attention has turned to other explanations based on radical alterations in structural organisation (including transformative structuralist proposals, saltation, alternative phenotypes in both time and space). This is supported by the observation that functionality is maintained even when structural organisation is drastically changed. In regard to the model research organism, the fruit fly (Drosophila spp.), it is remarkable how homeotic mutations (mutations that affect development) result in major alterations in adult flies, yet the flies remain functional:

172 

P. MCMAHON

[O]rganisms, considered as totalities, can sometimes resist disruption of their typical form caused by changes in genetic composition; [perhaps] they have powers of self-regulation against ‘internal’ (genetic) perturbation that are similar to those well-established powers of self-regulation against ‘external’ (experimental) perturbation that are possessed by most developing and some adult organisms. (Webster and Goodwin 1982)

It seems that processes at a higher level than gene expression (processes that are independent of, and not reliant on, genes) are responsible for the internal organisation at both the intra-cellular level and the extra-cellular arrangement and function of tissues and organs. Some evo-devo biologists question the process of gradual, cumulative selection as the only viable evolutionary mechanism (see West-Eberhard 2003, 471). They also question that genetic mutations are the only source of novelty. Alternative explanations look to phenotypic changes as the driver of change. Novelty need not depend on genetic mutations. The biologist T.H. Frazzetta (1975) talks of the “flexibility of integrated systems”. West-Eberhard views this in itself a basis for evolution as “the flexibility of systems may be a source of true innovation” (West-Eberhard 2003, 161). Notably the system as a whole undergoes transformation: Multidimensional adaptive plasticity accounts for coordinated change in several features at a time, and it explains why a sudden large change could be immediately adaptive or at least not disruptive. (Ibid.)

Radical Phenotypic Transformations von Baer’s Rule Under the modern synthesis it has been proposed that evolutionary genetic changes to developmental pathways would logically occur at the terminal points of developmental pathways. John Maynard-Smith proposed that only terminal stages of gene networks of development could be modified (by mutation) without deleterious consequences (see Fig. 8.1). The corollary of this principle is Karl von Baer’s proposal that specialist structures are added to a universal archetype (West-Eberhard 2003, 9). In developmental pathways, changes at earlier, fundamental stages would be more likely to be deleterious, while later changes (e.g., a specialised character) would be less likely to upset the more basic stages of development.

8  PHENOTYPE THEN GENE 

173

Fig. 8.1  A development pathway showing early and later stages. Changes to genes linked to earlier branches (unshaded) would have major and deleterious effects; therefore, modern evolutionary theory maintains only changes at the later stages (shaded) would be feasible—shown, in this case, at bottom right. (After John Maynard-Smith)

This is consistent with von Baer’s rule regarding embryogenesis—biological diversification occurs by branching out from a basic (unitary) plan into particular, specialised forms. John Maynard-Smith interprets this in terms of developmental pathways determined by gene networks with branching steps (Fig. 8.1). The earlier steps are fundamental and conserved, while the later branches allow some flexibility, and could be changed by genetic mutations followed by selection. This makes sense since changes at the outset the development of an embryo are likely to be disastrous, while later modifications may be accommodated more readily. This view is also consistent with Stuart Kauffman’s models of branching gene networks, leading to cell differentiation. However, early stages in ontogeny may not be as conserved as previously thought, as “we are now more aware that embryological characters themselves are extremely flexible, sometimes even arbitrarily so, and that exceptions to von Baer’s rule are rather frequent” (Willmer 2003, 35). Supporting this: “Profound changes during early ontogeny occurs in many animals, including mollusks, beetles, sea urchins, and fish” (West-­ Eberhard 2003, 10). Profound, early changes could account for radical

174 

P. MCMAHON

phenotypic change and support proposals of saltation as an evolutionary mechanism. The extended evolutionary synthesis (EES) puts emphasis on phenotypic plasticity. Phenotypes are not the fixed outcome of the central direction of genes: “[O]ntogeny is a condition-sensitive, bifurcating process that allows and even promotes polymodal adaptation” (ibid., 10). These proposals are supported by discoveries in molecular biology and genomics, that the same genes may be involved in different systems. Evo-­ devo, therefore, challenges conceptions such as von Baer’s rule, namely, that changes in early ontogeny are much less likely than at later stages.

Phenotypic Accommodation Just as literary evolution is based on a transformation of relations within the structure (the set of devices and elements that make up a text) leading to deformation (Chap. 6), biological evolution driven by phenotypic change can be explained by phenotypic accommodation, also called compensation by Frazzetta (1975). Accommodation occurs when an organism placed under environmental stress or a particular need adjusts its internal relations in response. This has been demonstrated in studies of the growth processes in crippled animals, for example, a goat born with missing front legs—as the kid grows the whole internal bone and muscle structure is changed enabling the goat to hop around with an upright posture (Slijper 1942). Phenotypic accommodation can be seen in the skulls of wild spotted hyenas. The bone structure has been modified from birth to accommodate the muscles needed to chew through bone (hyenas have one of the strongest bites in the animal kingdom). Much of the hyena’s diet consists of the left-overs (mainly bones) from the kills of other predators, such as lions. However, these skull bone structures are not present in infant hyenas. Nor do they develop in hyenas born in captivity, since they are fed with soft food. In wild hyenas, it is phenotypic accommodation that enables the consumption of hard materials, including hides and bones. The ‘Baldwin effect’ is another example; James Baldwin (1861–1934) was among the first evolutionary theorists to propose that phenotypic plasticity (with certain variant forms being more successful, especially in changed conditions) might lead to a particular variant of a phenotype surviving, and therefore being selected (Crispo 2007). These evo-devo ideas are not Lamarckian as the gene (and DNA coding) remain the most important hereditary mechanism. Through genetic accommodation,

8  PHENOTYPE THEN GENE 

175

phenotypic change can be encoded in the genome (invoking Waddington’s thesis of canalisation). A phenotypic change could be the result of a shift in the system to alternative modularities due to an environmental change or cue, or the accommodation of the phenotype itself to imposed stresses or changes. Many examples of this are found. A well-known one is an experiment reported by Twitty (1932): the eye of a salamander was transplanted into the embryo skull of a smaller species; the transplanted eye maintained its size but the host (the smaller salamander species) developed a larger optic cup to accommodate the larger eye. These changes may then be encoded genetically. In subsequent generations, genes that align with, are consistent with, and support the phenotypic change are selected—they are selected not because they determine the change but because they favour it (they are gene systems that have some alignment with the new phenotype). As the phenotype is selected, these genes follow. Only then does the adaptive phenotypic change become heritable. This process is known as genetic accommodation (see below). Even the famous homeobox (Hox) genes may have been be recruited according to Pat Willmer (Chap. 7). This is a divergent view to Darwinian adaptationism, in which new genetic variants (including those resulting from DNA mutations, but also meiotic recombination) are expressed providing a new phenotype, which is then subject to selection. Genetic Accommodation Evo-devo turns the tables on adaptationism, yet retains the gene (and genetic variability) as a key component of the phenotypic system. West-­ Eberhard has said that genes are followers rather than leaders: I consider genes followers, not leaders, in adaptive evolution. A very large body of evidence shows that phenotypic novelty is largely reorganizational rather than a product of innovative genes. Even if reorganization was initiated by a mutation, a gene of major effect on regulation, selection would lead to genetic accommodation, that is, genetic change that follows, and is directed by, the reorganized condition of the phenotype. (West-­ Eberhard 2005a)

This is explained through the concept of genetic accommodation. With genetic accommodation, the environment is brought into the equation: the consistent theme in evo-devo is that epigenetic factors (factors outside

176 

P. MCMAHON

of the genome) play an equally critical role in the development of a phenotype, and its heritable characteristics, as the gene (which after all in terms of DNA, RNA and associated systems is also hardware with its own epigenetic contributions). Genetic accommodation is a broader concept than genetic assimilation. Genetic assimilation, proposed by Conrad Waddington, shows how a phenotypic response to and environmental input is eventually encoded by genes, so that environmental sensitivity to that input is lost. Waddington illustrated this using the metaphor of an epigenetic ‘sloping landscape’ of valleys and ridges (or, in other words, discrete genetic systems), so that a marble rolled down the structure bifurcates into one valley or another. Waddington presumed that variant genetic systems are already in place, and that environmental cues result in a cascade along one ‘valley’ or another (a process he termed, canalisation). Thus, the marble rolling down the slope (representing the sequence of development steps) is diverted into one valley or another, depending on environmental conditions, However genetic accommodation is a broader concept than assimilation—rather than the new trait becoming genetically fixed, it may retain environmental sensitivity. Therefore, “[g]enetic assimilation is a special case of a more general phenomenon, called genetic accommodation” (Braendle and Flatt 2006). Underlying the notion of accommodation is the idea of alternative phenotypes—or at least systems that can be activated with an environmental stimulus. In the event of such a switch, the changed phenotype must accommodate the change. Thus, the first stage of genetic accommodation is the emergence of a novel phenotype (that may be already present cryptically or, in other words, as an inactive subsystem within the phenotype) and this change must be accommodated. The novel or reawakened system involves structural reorganisation and consequently stress—phenotypic accommodation is the organism’s ability to cope with this added stress. An adjustment occurs in the whole system or structure (the wider phenotype or the organism). This is possible due to the property of phenotypic plasticity. Table 8.1 outlines the main steps in the process of genetic accommodation. In the first step, a different phenotype emerges. What West-Eberhard (2005b) calls “innovative morphological form” may not be the outcome of a novel factor, such as a mutation or an environmental cue, but might already be present as an ancestral legacy. Therefore, it reflects past functionality (ibid.). This, taken together with the modular nature of the sub-­ systems that define alternative phenotypes, goes some way to explaining

8  PHENOTYPE THEN GENE 

177

Table 8.1  Steps towards evolution directed by the phenotype Requirements

Steps

 1. A variable population: a combination of phenotypic plasticity that is responsive to the environment and genetic variability  2. Developmental recombination: the organism consists of phenotypic systems that can recombine

Organisms are phenotypically changeable: they respond to environmental or genetic changes by altering the organisation of their phenotypic components Recombination of existing components occurs in response to the new input producing novel phenotypes; individuals vary in this response leading to variability among these types Those individuals with better reproductive success are preferentially selected under the new conditions; their phenotype spreads rapidly

 3. Selection: if the new form has a positive effect on fitness natural selection will favour it according to the changed conditions  4. Genetic accommodation: if the trait has a genetic component, then since the phenotype has been selected associated genes will follow

In cases where the new phenotypic trait has a genetic component, those genes are selected, increasing gene frequencies connected with the trait. The trait then becomes heritable (dependent on its associated genes)

After West-Eberhard (2005a)

recurrence (or atavism): the reappearance of ancient traits in more recent phylogenetic lines. The next step towards genetic accommodation in the novel phenotype is for it to become established in the population. It now represents a subpopulation, but under the new environmental parameters shows reproductive success, enabling it to spread through the population. This is natural selection at work. The stage of genetic accommodation comes next—some sources of phenotypic variation have a genetic component and, if so, selection for the phenotype will result in fixation of the associated genes. Critically, this is contingent on there being a genetic component to the phenotypic variation (West-Eberhard 2005a). Some variation is only phenotypic (outside of the genome)—evolution therefore depends on some kind of genetic component (or transmission to the next generation would not occur), but not all variation is associated with genes. Genes associated with the trait are then selected—in the case of genetic assimilation, the phenotype loses its environmental sensitivity in a process of assimilation (so the phenotype is expressed in the absence of the environmental cue, such as higher temperature).

178 

P. MCMAHON

However, genetic accommodation can also result in a phenotype that maintains its plasticity, as illustrated in Fig.  8.2. Evidence for

Fig. 8.2  Genetic accommodation: the phenotype (P) can be converted from form A into a new phenotype (B) by changing an environmental parameter (E), in this case by increasing temperature. (i) The slope of the line with the arrow indicates phenotypic plasticity; (ii) at the higher temperature conditions, the alternative phenotype (B) is selected replacing phenotype A; (iii) either the genes linked to trait B become fixed so now the original environmental condition (the lower temperature) no longer results in phenotype A, but maintains B and plasticity has been lost (1), a case of genetic assimilation, or plasticity is not lost and the lower temperature results in a reversion to phenotype A (2). (Adapted from Pigliucci et al. 2006)

8  PHENOTYPE THEN GENE 

179

genetic accommodation that maintains environmental sensitivity was reported by Suzuki and Nijhout (2006). These researchers showed that the larvae of a lepidopteran species, the tobacco hornworm (normally green) reverted to being black in mutant forms if exposed to lower temperatures. A related species displays colour polyphenism dependent on the ambient temperature (so larvae are black at 20oC and green at 28oC). Treated with heat shock, tobacco hornworm mutants also showed polyphenism—by selecting the green and black (monophetic) larvae over generations at the higher temperature, the green reverted to the black colour at the lower temperature: this study demonstrated the polyphenism was maintained in this line and was not genetically assimilated. Such instances where environmental sensitivity is maintained suggests that modular phenotypic traits could emerge under the right conditions or in phylogenetic sequences.

Internal Transformation and Deformation Geoffroy viewed transformism as a rearrangement of parts, with no new elements. This, as mentioned earlier, is a structuralist notion (Piaget 1971). As in structuralist linguistics, a system is not transformed by introducing new elements, but by changing the entire relational structure. Further, the structure of the organism is not dependent on its needs, its conditions of existence, but on internal structural adjustments. Organisms are transformed not by introducing new parts, but by their rearrangement—change could be substantial, with existing structures assuming new functions. The organism uses the internal resources it has at its disposal, rather than importing external or new elements. The German biologist, Carl Kielmeyer, proposed material resources available to organisms were limited, leading to compensation (as some structures are reduced to accommodate others) or preventing new structures emerging. This was rejected by Cuvier, who in his student days in Germany had learned methods of anatomical dissection from Kielmeyer. Structures, Cuvier suggested, were produced as needed—presumably through divine intervention. The autonomy of an evolutionary process in literature, and its jumpy nature, has themes in common with early twentieth-century attempts by biologists to account for macroevolution. Common features of literary formalism (especially in later Russian theory and the structuralist ideas developed in Prague) with earlier French ideas on biological transformism

180 

P. MCMAHON

are apparent: types are homologous because their fundamental structure (the links between component parts) remains intact, while the structural arrangement of individual parts is maintained but individual parts assume different functions through transformation. Ear bones of the mammalian inner ear can be identified in the jaw bones of reptiles, for example. Fewer hyoid bones in humans than cats suggest some have changed functions in parts of the skull. Just as Russian formalists envisaged literary change to be sudden, entailing an internal transformation and rearrangement of internal forms, Geoffroy’s saltation theory proposed that species emerge suddenly due to internal adjustments to environmental pressure. Geoffroy was interested in teratology: monstrosities demonstrated a seemingly radical change yet retaining previous structural arrangements. The important point to him was that change occurred within the framework of an existing arrangement—to Geoffroy this was the Principle of Connection. Now there is a renewed interest in the hopeful monster (Theißen 2006)—the radical evolutionary hypothesis suggested by Richard Goldschmidt. Basic to the formalism of Geoffroy and others in the early nineteenth century, and the return to it at the turn of the twentieth century by Driesch, D’Arcy Thompson, and others, is that the organism acts as a whole, it is not transformed part by part, but by a rearrangement in the whole structure. Again, we can see a parallel with later Russian formalism: Tynianov insisted that literature can evolve, not through external social conditions or circumstances but within the scope of literature itself. Change involved the rearrangement of existing forms. He did not reject the possibility of external modification and contextual influences (including socialism) but distinguished this type of modification from evolution. In ‘On literary evolution’, he states such “major social factors … replace the study of evolution of literature with the study of modification of literary works” (Tynianov 2019). Deformation as a driver of literary change and evolution results from a radical change in the relations between structural forms as proposed by the Russian formalists and their Prague successors. As devices assume dominance, other elements become ‘deformed’—there is a strain put on the system. Eventually the tension is resolved, and forms assume equivalence. This leads to a new constructive element (a form that assumes dominance in a new set of relations) suddenly replacing the old one. To the formalists, such a process was necessary in literary change, in order to defamiliarise. It enabled ‘actualisation’, the dialectic response to

8  PHENOTYPE THEN GENE 

181

‘automatisation’, in which literariness was replaced by practical (realist or ‘motivated’) text (Galan 1979). Therefore, literature changes according to a continuing dialectic relation between actualisation and automatisation. As discussed later, realist literature is more susceptible to (even a product of) ideology (Chap. 9). On the contrary, literariness (actualised literature) constitutes an autonomous state resistant to ideology. Parallels are found between this formalist version of literary evolution and radical phenotypes as possible drivers in biological evolution. Various ideas have been broached, one of the earliest being Geoffroy’s hypothesis that teratological changes in embryos resulted in the formation of new types or species (saltation), mutationism and the ‘hopeful monster’ hypothesis of Richard Goldschmidt, a century later. Presently, the importance of modularity is becoming evident. This can occur in different contexts: “Modular organization in biological systems results from patterns of differential integration that exists within and among units defined in genetic, developmental, functional, or evolutionary contexts” (Gerber and Hopkins 2011). Discrete, complex sub-systems and structures that are semi-isolated can be combined into novel forms. Ideas have been put forward that the combination of traits in novel formations could also underly speciation. A Revival of Saltation Theißen (2006, 2009) proposes that saltation as an evolutionary mechanism, which has been dismissed by the modern synthesis should be reconsidered—although likely to be rare it could explain novel forms in macroevolution (while Darwinian natural selection is more important to microevolution, particularly in populations)—particularly promising are possible homeobox regulatory gene mutations. However, to propose saltation (or major phenotypic change) as a possible mechanism of sympatric speciation is challenging to a gradualist account of evolution: Ever since Darwin, there has been a tension between selectionists and developmentalists over the question of gradualism versus saltation and selection versus variation. Does form evolve by a series of small modifications, each one mediated by selection? Or does complex change in form originate suddenly due to a developmental change? (West-Eberhard 2003, 471)

182 

P. MCMAHON

Proposals that saltation (sudden change) contributes to macroevolution have been marginalised or considered ‘mad’ by proponents of the modern synthesis of biology. Stephen Jay Gould, in an essay on the critical response to Richard Goldschmidt’s ideas, regarded this as short-sighted. Gould was particularly interested in heterochrony, in which very different forms occupy the same life cycle. Some larvae (e.g., the motile larvae of sponges and sea urchins) have structures more complex than the adult, sessile forms. Based on such observations, some biologists now take more seriously proposals that sudden phenotypic changes in structural arrangements may be evolutionary mechanisms. In a similar way to the literary evolution model of Tynianov, an evo-devo hypothesis would suggest that the whole structure is re-arranged, and the function of constituent elements may change. In evo-devo biology, anomalies are not excluded as potential evolutionary mechanisms: “Rare phenotypes (anomalies) are valuable and unjustly neglected indicators of trait origins, for recurrent anomalies are material for selection and evolution in new directions” (West-Eberhard 2003, 197). Furthermore, variation need not be small or incremental, but might entail a substantial change in the phenotype: “[L]arge discrete variants are not always devastating to survival and reproduction, but can be bearable handicaps that eventually prove advantageous and become widespread established traits” (ibid., 374). Also, modularity may function to cushion the effects of a major phenotypic change, thus reducing disruption: “Thanks to the modular nature of phenotypic development, innovative variants are semi-isolated from other functioning traits, and therefore not so disruptive as they could be if all development were interconnected” (ibid.). The case of a goat deformed at birth that adopted a bipedal gait, with massive structural changes in the pelvic bone structures is a case in point (Slijper 1942). Such examples suggest phenotypes can accommodate more severe changes than envisaged by many, suggesting a mechanism for evolution: Two-legged goats and other flexible ‘monsters’ indicate that the outer limits of adaptive plasticity—the location of constraints—are less severe than may be suggested by the stereotypy of normal development and physiological homeostasis. (West-Eberhard 2003, 297)

Could such an event have led to bipedalism in humans? it has been asked—however, it has also been pointed out that these changes were not demonstrated to be heritable.

8  PHENOTYPE THEN GENE 

183

The lack of direction in such an evolutionary scenario and the possibility of recurrence (as old forms are recouped by more recent phylogenies) has similarities to the formalist account of literary evolution developed in Russia. In both cases, the old is mixed with the new, changes can be sudden and radical, and there is no ‘telos’ or an end-goal in the evolutionary process. Evolution seems more cyclical than linear, and represents an ongoing transformative process, rather than a pre-defined purposefulness towards a goal (such as greater complexity). In fact, in biological phylogenies, less complex sub-systems may recur.

Is Homoplasy Really an Instance of Recurrence? Convergence in evolution needs to be differentiated from homology—in convergence similar structures evolve independently in separate phylogenies, while homological structures are related by common descent. The common explanation for convergent evolution is that under similar environmental pressures, selection works over time to produce very similar structures in unrelated organisms. A well-known instance of convergence, also known as homoplasy, has occurred with the evolution of the eye in cephalopods and vertebrates (Fig.  8.3). Camera eyes are considered superior to compound eyes (found in crabs and insects, for example). As pointed out in Chap. 2, Georges Cuvier provided evidence that the two groups (cephalopods and vertebrates) had nothing in common in their internal structural arrangements and could be considered non-­homological. Currently biologists are still in agreement that the two phylogenies are distant (anatomically and according to molecular sequences) and the two taxa cannot be considered homological. However, both have developed the camera eye independently, a striking example of evolutionary convergence (Conway Morris 2003, 151). Simon  Conway Morris proposes that evolution takes place along pre-existing pathways or ‘adaptive landscapes’, leading to an increase in biological complexity. Similar proposals were made in the nineteenth century by Lamarck, Haeckeland others, who endorsed orthogenesis (Chap. 7). Convergence is therefore not viewed as a chance occurrence, but as the result of pre-set pathways. Under this view, there are far fewer evolutionary options than would be expected from the selection of chance mutations, the position of the modern synthesis. Sticking with the octopus, another interesting convergent feature is the autonomous forms of behaviour found in octopi, including apparent playfulness. Playfulness in species radically removed from higher vertebrates is not generally observed. Numerous examples of homoplasy are reported in

184 

P. MCMAHON

Fig. 8.3  Camera eyes have evolved independently: the vertebrate eye (left) and the octopus eye (right). (1) Retina; (2) Neurones; (3) Optic nerve; (4) Blind spot (in vertebrates). These are not homological structures and the two taxonomic groups are unrelated. (Wikimedia, Creative Commons)

Conway Morris’s book whereby unrelated taxa have evolved similar structures—one example is halteres that have evolved in two separate groups of insects—these drumstick-like structures are sensory adaptations for balance and are modified from hind wings in flies and forewings in streptiserans, with a similar function. They are not homological, but convergent (structures that have evolved in different taxa to perform a similar function). Pat Willmer (2003) raises unsolved mysteries that imply convergence. Do larval forms, for example, reflect their mode of existence (whether they are sessile and ciliate, or motile, for example?). Are these forms an effect of their environment? Similar larval forms in very different phylogenies would suggest this. Or could convergence be an outcome of modularity, whereby traits reappear in phylogenies during evolution leading to an appearance of independently evolved structures? It is widely accepted that organisms have a modular structure. Modules (discrete subsystems) may be cryptic, only to reemerge under certain conditions. Therefore, some examples of homoplasy could turn out to be homologies (West-Eberhard 2003). Certain environmental conditions could trigger the activation of previously inactive modules. Structuralist models propose that novel structuralist arrangements may give rise to new forms, with similarities to formalist notions of literary evolution. While not directed (i.e., orthogenetic) the choices may be constrained since change depends on the recurrence of previous forms. This is

8  PHENOTYPE THEN GENE 

185

especially the case when new elements are not introduced into the system. This structuralist account in both literary and biological evolution suggests that available forms are more restricted than in the neo-Darwinian explanation (where the potential for mutational changes are apparently limitless). Therefore, homoplasy may reflect a restricted range of possible (functional) structures that emerge in different phylogenetic lineages in evolution.

References Bolker, Jessica A. 2000. Modularity in Development and Why It Matters to Evo-­ Devo. American Zoologist 40 (5): 770–776. https://doi.org/10.1668/0003-­ 1569(2000)040[0770:MIDAWI]2.0.CO;2. Braendle, Christian, and Thomas Flatt. 2006. A Role for Genetic Accommodation in Evolution? BioEssays 28 (9): 868–873. https://doi.org/10.1002/bies.20456. Conway Morris, S. 2003. Life’s solution: inevitable humans in a lonely universe. Cambridge, UK: Cambridge UP. Crispo, Erika. 2007. The Baldwin Effect and Genetic Assimilation: Revisiting Two Mechanisms of Evolutionary Change Mediated by Phenotypic Plasticity. Evolution 61 (11): 2469–2479. https://doi.org/10.1111/j.1558-­5646.2007. 00203.x. Dawkins, R. 1991. The Blind Watchmaker. 2006 edition. London: Penguin. Eldredge, N. and Gould, S.J. 1972. Punctuated equilibria: an alternative to phyletic gradualism. In Schopf, T.J.M. (ed.). Models in Paleobiology. San Francisco, CA: Freeman Cooper. pp. 82–115. Frazzetta, T.H. 1975. Complex Adaptations in Evolving Populations. Sunderland, MA: Sinauer. Galan, F.W. 1979. Literary System and Systemic Change: The Prague School Theory of Literary History, 1928–48. Proceedings of the Modern Language Association (PMLA) 94 (2): 275–285. Gerber, Sylvain, and Melanie J.  Hopkins. 2011. Mosaic Heterochrony and Evolutionary Modularity: The Trilobite Genus Zacanthopsis as a Case Study. Evolution 65 (11): 3241–3252. https://doi.org/10.1111/j.1558-­5646.2011. 01363.x. Gould, S.J. 1977. Ontogeny and phylogeny. Cambridge, MA and London, UK: Belknap Press (Harvard UP). Hall, Brian K. 2003. Evo-Devo: Evolutionary Developmental Mechanisms. The International Journal of Developmental Biology 47 (7–8): 491–495. https:// doi.org/10.1387/ijdb.14756324. Larsen, Ellen. 2003. Genes, Cell Behavior and the Evolution of Form. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 119–132. Cambridge, MA: MIT Press.

186 

P. MCMAHON

Nanjundiah, Vidyanand. 2003. Phenotypic Plasticity and Evolution by Genetic Assimilation. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 245–264. Cambridge, MA: MIT Press. Piaget, J. 1971. Structuralism. London: Routledge & Kegan Paul. Pigliucci, M., C.J. Murren, and C.D. Schlichting. 2006. Phenotypic Plasticity and Evolution by Genetic Assimilation. The Journal of Experimental Zoology 209 (Pt 12): 2362–2367. https://doi.org/10.1242/jeb.02070. Raff, R.A. 1996. The Shape of Life. Chicago: Chicago University Press. Slijper, E.J. 1942. Biologic-Anatomical Investigations on the Bipedal Gait and Upright Posture in Mammals, with Special Reference to a Little Goat, Born without Forelegs. I. Proceedings of the Koninklijke Nederlandse Akademie Wetenschappen 45: 288–295. Suzuki, Y., and H.F.  Nijhout. 2006. Evolution of a Polyphenism by Genetic Accommodation. Science 311 (5761): 650–652. https://doi.org/10.1126/ science.1118888. Theißen, Günter. 2006. The Proper Place of Hopeful Monsters in Evolutionary Biology. Theory in Biosciences 124: 349–369. ———. 2009. Saltational Evolution: Hopeful Monsters are Here to Stay. Theory in Biosciences 128 (1): 43–51. https://doi.org/10.1007/s12064-­009-­0058-­z. Twitty, V.C. 1932. Influence of the Eye on the Growth of its Associated Structures, Studied by Means of Heteroplastic Transplantations. Journal of Experimental Zoology 61: 333–375. Tynianov, Jurii. 2019. On Literary Evolution. In Permanent Evolution: Selected Essays on Literature, Theory and Film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 267–282. Boston: Academic Studies Press. Venditti, Chris, Andrew Meade, and Mark Pagel. 2009. Phylogenies Reveal New Interpretation of Speciation and the Red Queen. Nature 463: 349–352. https://doi.org/10.1038/nature08630. Webster, G., and B.C.  Goodwin. 1982. The Origin of Species: A Structuralist Approach. Journal of Social and Biological Structures 5 (1): 15–47. https:// doi.org/10.1016/S0140-­1750(82)91390-­2. West-Eberhard, M.J. 2003. Developmental Plasticity and Evolution. Oxford University Press. ———. 2005a. Developmental Plasticity and the Origin of Species Differences. Proceedings of the National Academy of Sciences of the United States of America 102 (Suppl 1): 6543–6549. https://doi.org/10.1073/pnas.0501844102. ———. 2005b. Phenotypic Accommodation: Adaptive Innovation Due to Developmental Plasticity. Journal of Experimental Zoology Part B: Molecular and Developmental Evolution 304 (6): 610–618. https://doi.org/10.1002/ jez.b.21071. Willmer, Pat. 2003. Convergence and Homoplasy. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 33–49. Cambridge, MA: MIT Press.

CHAPTER 9

The Interpreting Organism

The Affective Fallacy Formalist literary theory in the first half of the twentieth century downplayed factors outside of the text, attributing meaning to relations within the literary object. Formalists on both sides of the Atlantic took an objective (and scientific) approach to text. The New Critics dismissed the role of the author in textual meaning: meaning was bound up in the text itself, an object that could be subjected to analysis. The Russian formalists recognised authorial input, yet downplayed the notion that a text reflected the psychology or background of the author. In both formalist movements, expressive realism, prominent in the nineteenth century was rejected. In ‘The Intentional Fallacy’, the New Critics William C. Wimsatt and Monroe Beardsley rejected authorial intention as a locus of meaning (Chap. 4). They also rejected the role of the reader in attributing meaning in a second essay, ‘The Affective Fallacy’. In the same fashion as their first essay, the authors propose that meaning is to be found in the text itself, its device, its oppositions and paradoxes. The reader, his/her background, belief systems, or historical situation, has little input into textual meaning. Rather, meaning needs to be elucidated by objectively analysing the text on its own merit. They criticised the subjectivity of the reader, which creates “a confusion between the poem and its results (what it is and what it does)” (Wimsatt and Beardsley 1949). To the New Critics, the key was the © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_9

187

188 

P. MCMAHON

expectation of unity in text (Easthope 1991, 21). In this sense, the New Critics followed the Romantic tradition. Coleridge suggested that a poem proposes to itself “such delight from the whole, as if compatible with a distinct gratification from each component part” (quoted in ibid., 21). The waning of New Criticism in the late twentieth century is associated with an increasing focus in literary theory on the role of the reader. In fact, Wimsatt’s student, E.D.  Hirsch, exemplified this break from the New Critics’ stance, by moving towards an emphasis on the author and the context of a literary work: “Thinking in part of Wimsatt, Hirsch insisted that appealing to intention is the only thing you can do in order to establish … ‘validity in interpretation’” (Fry 2012, 42). In order to paraphrase a text, for example (something insisted on in educational institutions to avoid plagiarism), the reader or student has to come to some conclusion about the author’s intended ‘meaning’ (ibid., 42). In modern literary theory, the reader takes a prior position in the construction of meaning. Reader-response theories have criticised the notion of meaning embodied by a text. Meaning is much more subjective than this and comes about through an interpretative process. Hence, replacing and superseding formalist ideas is a focus on interpretation or hermeneutics. Philosophical positions influenced by Heidegger and Gadamer take hold. The shift to the reader reflects the movement from structuralism to a reader-centred poststructuralism, as the latter insists the reader is active in producing textual meaning. This is a significant divergence from the premises of New Criticism, especially ‘the affective fallacy’ and its dismissals of the role of the recipient and reader. Structuralism and formalism have buckled in the face of Jacques Derrida’s critiques (which, according to some, represent a re-evaluation of Husserl’s phenomenology, the subject of Derrida’s doctoral thesis), as well as the identification of textual meaning with the reader. An emphasis on interpretation in reader-response theories has largely supplanted the literary formalism of the early- to mid-twentieth century. In parallel with the rise of poststructuralism (breaking away from structuralism and its formalist heritage), this represents a trend, not only in literary theory but also in ideas related to biology, notably biosemiotics. Biosemiotics turns to phenomenology. Taking up the views of Heidegger, Gadamer and other philosophers, to biosemioticians organisms do not exist first and then interpret after: they are already interpreting beings and they cannot be divorced from their external conditions. And like the proverbial reader, they are subjects that produce meaning.

9  THE INTERPRETING ORGANISM 

189

Biosemiotics Alternative schools of thought in biology have questioned the primary role of structure endorsed by early biologists, changing the focus onto how organisms communicate with and interpret external factors (the environment, other organisms). This has instigated the discipline of biosemiotics, which has become increasingly influential. Biosemiotics emphasises relations of the organism with its environment. Biosemiotics is based on theories of interpretation. One of its early protagonists, Anton Markoš, suggested the name biohermeneutics for the growing discipline (Barbieri 2019). Ideas of holism expressed by thinkers such as Hans Jonas, Susanne Langer, Francisco Varela, Humberto Maturana and Jakob von Uexkull have been developed by biosemioticians. The organism is seen as a subject, a subject of its environment and mode of living. More than this, it engages in active realisation. Organisms have their own phenomenological experience. This means they have qualitative diversity, and biology should be approached using both qualitative and quantitative methods. Biosemiotics is critical of behavioural/genetic explanations and considers the modern synthesis of Darwinian theory and genetics as non-semiotic. It is adamantly non-reductionist, rejecting the assumption that biological phenomena can be reduced to physical processes. The ‘Umwelten’ and Biosemiotics Readers bring different interpretations to any single text and, therefore, meaning becomes subjective. Is the experience of the biological organism also meaningful? The formalist approach of Cuvier, followed by others, is to analyse the organism as object and elucidate how its internal relations enable it to function within its conditions of existence. The biologist, Jakob von Uexkull (1864–1944), born in Estonia to an aristocratic German family, rejected the possibility of an objective knowledge of general organism-environment relations. In von Uexkull’s scheme, each organism is a ‘subject’ and has its own species-specific environment, the Umwelt. Umwelt is differentiated from the general surroundings, the environment as a whole (Umgebung) (Sullivan and Malkmus 2016). In fact, von Uexkull claimed humans have their own umwelt and that an understanding of the environment in its totality is impossible. This led to disagreements with his colleague, Ernst Cassirer, who attributed to

190 

P. MCMAHON

humans alone the ability to achieve objective knowledge, while conceding other animals could only have subjective experiences. ‘Environment’ to von Uexkull meant an endless series of linked umwelten; von Uexkull interpreted the umwelt in cybernetic terms (Fig.  9.1). Each umwelt could be seen as feedback loop between the organism and its particular needs. For example, the female mosquito of many species requires blood to produce and lay eggs. The umwelt of the mosquito consists of carbon dioxide (which helps it locate hosts), the odour of chemicals such as lactic acid and octanol and an area of skin lacking hair to enable it to insert its proboscis. These salient factors are meaningful in a species-specific sense. The sensory signals initiate the mosquito’s active response (the ‘effect’ of inserting its proboscis to take up blood) completing the ‘functional circle’. Thus, the ‘environment’ for a particular species becomes extremely restricted, limited only to the needs of the species to survive. Jakob von Uexkull’s umwelt theory is phenomenological, rather than structuralist, although he was influenced by Saussurean linguistics and the Tartu structuralists, including Yuri Lotman. Umwelt describes the environment as experienced (and interpreted) by the individual and thus has a

Fig. 9.1  Umwelt as a cybernetic system (public domain, University of Hamburg)

9  THE INTERPRETING ORGANISM 

191

firm phenomenological basis. Phenomenology in general investigates the role of intention and interpretation and, therefore, its premises can be applied to animals (such as the female mosquito, which has a clear intention to find blood): “von Uexkull’s umwelt theory is in effect a biological theory that aims to explain how it is that (many) animals too have phenomenal worlds, and how these differ from species to species and from one individual to another” (Tønnessen et al. 2018). There must be a minimal umwelt, as Tonnessen suggests: “It is arguably the case that any organism endowed with an Umwelt relates to something specific in its environment as nutrition, by foraging and consuming food, and that it relates to something specific as its medium, by navigating in its environment. This implies that the minimal Umwelt is constituted by the functional cycle of nutrition and the functional cycle of the medium” (Tønnessen 2022). Tonnessen and colleagues point out this phenomenal aspect does not apply to other living things (plants and fungi, for example) but that these organisms also relate to their environment with signs: [W]e acknowledge that all Umwelt experience, whether it is of a phenomenal nature or not, is sign-based (and that even organisms that are not endowed with an Umwelt relate to their environment in a sign-based manner). (Tønnessen et al. 2018)

This is a reference to the field of biosemiotics, which is partially based on von Uexkull’s ideas. An emerging field in alternative approaches to biology, biosemiotics engages with sign systems, following Peircean semiotics (and his adage, ‘a sign is something that stands to somebody for something’). A definition of biosemiotics is provided by Jesper Hoffmeyer: Biosemiotics is the name of an interdisciplinary scientific project that is based on the recognition that life is fundamentally grounded in semiotic processes. (Hoffmeyer 2008, 3)

First called zoosemiotics, as the focus was on the way cognisant animals relate to their particular environment (the umwelt), the discipline of biosemiotics now encompasses all living things and their sign systems. The organism interprets its environment based on sign systems. The organism connects to its inorganic environment, as well as other organisms—in this environment-world or phenomenon-world the organism is a subject. The world is imbued in signification—our biology classes may have used the

192 

P. MCMAHON

metaphors: signal, message, code—but these are considered real in biosemiotics. This is one reason biosemioticians reject the reduction of biological processes to those defined by chemistry and physics. The strength of biosemiotics is that it introduces the contextual and subjective existence of the organism. A weakness of formalist approaches (such as the analysis New Critics apply to texts) is that external factors, the context in which a text is produced or read, are disregarded. Biosemiotics is perhaps the up-to-date approach to an alternative biology. Context is everything—objects do not exist in isolation but only in relations with other objects. Organisms also exist (and function) in context, not as isolated objects. The organism and its niche cannot be considered separately—each defines the other—drawing inspiration from the Umwelt model of Jacob von Uexkull. Organisms interpret the world in a species-specific manner. Von Uexkull reminds that life does not exist outside of its environment and the environment creates life. This is especially relevant to the effect of human practices on ecology as we take species “out of their original biocoenosis” (Kull 1998). Removal from connections with other species gives rise to weed and other problems. Kalevi Kull criticises the notion that humans can replicate nature—human signs cannot copy all the details of non-human signs and endeavours to do so is a process of simplification. The complexity of nature (e.g., elucidated in the work of Michel Serres) is beyond our understanding (Conley 1997). Some of these ideas are becoming increasingly relevant with the development of concepts such as the holobiont and metagenome (Simon et al. 2019). Biological organisms encounter the world through an indirect mediation (Weber 2002). They live in an experiential world. Expressiveness is just as vital as functional adaptation. Certain aspects of the environment are meaningful to the organism. This meaningfulness is ignored by behaviourist and computer-simulated models. In relation to linguistics, a computer may be able to manipulate elements (or the syntax) but cannot grasp their meaning. In meaning-making by an organism the object becomes represented internally, the ‘innerwelt’, followed by a response when action is taken. The emphasis in biosemiotics on communication is in keeping with current ideas. Organisms survive, even exist, because they receive and send information. They interpret their environment. They detect changes and respond accordingly; they communicate with other organisms—their world is a network of interactions. They are embedded in this network and

9  THE INTERPRETING ORGANISM 

193

cannot be considered as isolated entities. As such, biosemiotics is antireductionist. However, biosemiotics differs from structuralist approaches in major ways. The Prague Linguistics Circle (Chap. 6), despite elucidating functional characteristics common to texts (such as the representative function), placed structure ahead of function. Roman Jacobson stuck to Saussure’s dyadic structure (Chandler 2007). Biosemiotics, on the other hand, turns to the triadic sign of Charles Peirce. Based on Peirce’s philosophy of the sign, biosemiotics accepts that external reality can be grasped to a certain extent—that biological processes begin with real objects, albeit caught up in contextual relationships. Structuralism puts relations within the structure first—objects are delineated according to these relations. Network theory can better explain the biosemiotics world than structuralism. Biosemiotics expands the organicism of Kant, Coleridge and Cuvier (and New Criticism) to a more extended view of the organism—one that merges with its environment in a welter of signs. This is reminiscent of a dialectical biology in Roger Lewontin’s account—where the organism and environment are in constant communication, or a dialectic relation, so that a niche should be considered as a unit (Lewontin 2000). In addition, in common with biosemiotics, the organism constructs its own external conditions. On a planetary scale, this view is perfectly valid: the composition of gases in the atmosphere is a biological phenomenon—the earth’s atmosphere (its current composition) has been created largely by oxygen-­ exchanging biota, with the older anaerobes surviving only where oxygen has not penetrated. The rainforests and cyanobacteria of the oceans are considered the ‘lungs of the planet’ for this reason. Life regulates its own planetary conditions. James Lovelock devised the ‘black and white daisy’ model to illustrate the homeostatic regulation of temperature on earth by life: as populations of black daisies increase, the sun’s radiation is absorbed heating the earth to a critical point above which the daisies cannot survive. But the white daisies which reflect the sun’s heat survive better than the black daisies, and as the population of white daisies increases, more heat is reflected, and the earth cools. The carbon dioxide composition of the atmosphere is not coincidental or irrelevant to global temperatures according to Lovelock as has become clear with climate change. But niche construction applies at the species level too—the beaver constructs its dam, and the earthworm, ancestrally an aquatic organism, liquifies soil. Such observations of niche

194 

P. MCMAHON

construction support the umwelt concept: that the organism cannot be separated from its environment and, furthermore, is active in creating its environment. Differentiating Coding and Peircean Semiotics While Thomas Sebeok (Fig. 9.2), a founder of the discipline of biosemiotics, at first restricted his semiotics to the animal kingdom, in the 1980s he expanded ‘zoosemiotics’ to include all living creatures (Barbieri 2013). The remit of biosemiotics was very broad. The Belgian biochemist, Marcel Florkin applied Saussurean semiotics to the molecular code. Marcello Barbieri identifies two versions of biosemiotics, Peircean and coding versions. The latter he claims is not dependent on interpretation. To Peirce however, ‘the universe is perfused with signs’ and an interpretant is always needed. Sebeok and others followed the Peircean idea and applied this principle to all living systems. This is rejected by Barbieri. He reminds that only higher organisms have the capacity for representation and that this is an internal process. Barbieri disagrees that all biological signalling comes under the Peircean sign (in particular the idea of the extended universe of signs) and distinguishes organic code from what he terms Peircean biosemiotics (Barbieri 2019). Organic code is meaningful, but it is arbitrary and mechanistic, and meaning does not arise from interpretation, he argues. He rejects the premises of Naturphilosophie idealism (Goethe, von Baer and others) that

Fig. 9.2  Thomas Sebeok with his parents in Hungary c.1924 (public domain)

9  THE INTERPRETING ORGANISM 

195

lies behind much of biosemiotic thought, endorsing instead mechanism as the best explanation of coding. According to Barbieri’s separation of code from interpretation, only animals with a certain level of neurological development can interpret their surroundings. They have what Barbieri calls an interpretive brain. A snake will not pursue a prey that hides, but a higher vertebrate will. ‘Lower’ animals like lizards and snakes are responsive due to an instinctive brain (they do not interpret environmental cues but respond instinctively to them). And it follows that organisms with no nervous system (bacteria, for example) respond through organic codes— not interpretation, which is proposed by adherents to biosemiotics. Here, Barbieri parts ways from his biosemiotics colleagues. He broke from the school founding his own group, Code Biology. Divergence from Structuralism Biosemiotics was initially influenced by structuralist ideas (e.g., Yuri Lotman equated the ‘innerwelt’ with Saussure’s langue) but currently biosemiotics is based on the Peircean triadic sign, rather than the dyadic signifier-signified. In philosophical terms, it  is closely aligned with both hermeneutics, including the ideas of Heidegger and Gadamer, and pragmatism. A functionalism underlies biosemiotics, as the organism’s relations to its environment are based on functional needs. Biosemiotics takes up Peirce’s triadic system, which consists of a sign, an object and an interpretant. This triad acts as a unit—its components cannot operate in pairs. “Nothing is a sign unless it is interpreted as a sign” (Peirce 1960). Peircean ideas on pragmatism are adopted in place of structuralist ones: the organism (and the human) interprets (reads) objects in the environment and, therefore, engages with the referent. In other words, meaning does not lie with the inner relations of the organism but with continuous and extended relations with real objects in the environment. In contrast to organic codes (Barbieri 2019), meaning is interpreted through the Peircean interpretant. Peirce’s abduction has this in common with interpretation: a conclusion is reached from limited data—the basis of his pragmatism. Interpretation is therefore a jumping to conclusion based on the data provided—usually for practical purposes this works well. Charles Peirce stated that a sign is something that stands for something. To structuralists (at least in what might be considered ‘pure’ versions) there is no such ‘something’ or referential object. Therefore, Louis Helmsjlev’s structuralist linguistics is inconsistent with Peirce’s sign

196 

P. MCMAHON

system. Sharing common ground with Noam Chomsky’s generative grammar, in this regard language is seen foremost as an expression, not a means of communication. Language is viewed by Helmsjlev as immanent to structure so that linguistic facts needed to be separated from the non-­ linguistic. Jacques Derrida also rejected the triadic structure, sign-­ interpretant-­ object. However, biosemiotics also parts from the more communicative interpretation of Saussure’s ideas, Prague structuralism. As discussed in Chap. 6, the Prague circle of linguists retained Saussure’s system, including its paradigmatic and syntagmatic poles (interpreted as metaphor and metonym). A revised form of structuralism, (‘after-­ structuralism’, a poststructuralist development that encompasses the reader) provides an alternative model to the semiotic one. Notably, rather than the Peircean sign it adheres to Saussure’s dyadic signifier/signified pairing (Chandler 2007, 34).

A Revised Structuralism In literary theory, reader-response theory has largely supplanted formalism. A phenomenological approach is taken, which has also informed poststructuralist theory. However, Lacanian-inspired literary theory developed in Britain a few decades ago retained the principles of structuralism—according to this school of thought, poststructuralism is defined by a shift of emphasis from text to reader. A method of poststructuralist literary analysis, developed in Britain in the 1970s and 80s, took up Lacanian psychoanalysis (see below, Ideology and Realism). This school, which was also influenced by the scientific Marxism of Louis Althusser, proposed that the process of signification, operating in the symbolic order, structures conscious thought. Literary theorists applied Lacanian ideas to the interaction of literature and the reader. The unity embodied in the New Critics’ text rejected context and external factors. But poststructuralist literary theorists point out that behind this unity is both an implied author and a meaning that could be changed each time it is read: “If the text could no longer be treated as a complete, self-sufficient object, then the applied empiricism presuming the text was simply there outside any theory and practice for its construction had to go” (Easthope 1991, 19). A Lacanian poststructuralist text is decentred, in that reading leads to multiple interpretations. While signifiers have a “physical identity fixed on the page.. the signifieds are not fixed and cannot so be fixed. Any text especially one such as a poem, is constantly read and re-read in different ways—by different

9  THE INTERPRETING ORGANISM 

197

people, by the same people at different times in their lives, by different people at different periods in history. [Therefore], the meaning of a text is always produced in a process of reading” (Easthope 2002, 7 italics original). According to Antony Easthope, the shift in emphasis from texts to the reader was definitive to the structuralist revision into a form of poststructuralism: “Structuralism becomes transformed to post-structuralism when the structures of text are seen to be always in and for a subject (reader and critic). The text of structuralism is intransitive, that of post-structuralism transitive” (Easthope 1988: 33). The subject is brought into play but as occupying positions that are not finally defined by the text (or structure). Lacanian-inspired literary theory demonstrated how texts in a circular interactive process position us as subjects. The subject is dispersed but its unconscious production of meaning can be discerned in the text itself: “Possibilities of meaning … circulate between text, ideology and readers whose subjectivity is discursively constructed and displaced across a range of knowledges” (Belsey 2002, 139). This statement by Catherine Belsey is indicative of the Althusserian heritage of British poststructuralism and its focus on the ideological function of texts. However, it also shows that theory had already shifted from an Althusserian focus to a poststructuralist one where the reader entered into a text’s subjectivity. Antony Easthope (1988) jokingly suggested the term ‘after-­ structuralism’ was more apt for this movement since it represents a revision of the high structuralism of the 1960s, rather than its complete supplantation. In this sense, it differs from postmodernism and the theory that informs it (also frequently referred to as ‘poststructuralism’). Easthope’s final publication criticised these postmodernist versions of poststructuralism as ‘privileging difference’ (Easthope and Belsey 2002), and as embodying a distinct move away from Saussurean linguistics, as well as ideas emerging from practitioners of psychoanalysis. For the Saussurean/Lacanian system depends not on an ‘originary’ difference (the favoured postmodernist position) but on what Peter Dews calls an interplay of ‘identity and difference’ (Dews 1987). Although Jacques Derrida’s famous critique of structuralism in his paper ‘Structure, sign and play’, presented at John Hopkins University in 1966, launched theory into poststructuralist forms, including deconstruction and a kind of postmodernist phenomenology (Lash 2002), the revised structuralist ideas (which to Antony Easthope represented a transition to poststructuralism) turned to Lacanian psychoanalysis (Belsey 2002;

198 

P. MCMAHON

Easthope 1988). The static structure was reformulated as a process; the process of signification. Jakobson’s opposing poles of metaphor and metonym informed this change: in some ways the Prague Linguistics Circle prepared the ground for the transition to ‘after-structuralism’, as Easthope put it. Prague theorists, as discussed earlier (Chap. 6), rendered Saussure’s strict separation of synchrony and diachrony in linguistic theory into a systemic model of language as a transformative structure. In addition, the speaking subject assumed a new prominence thereby “enabling the Prague scholars to heal the other Saussurean rift between the langue – or the system – and the parole – or the individual utterance” (Galan 1985). Although the ‘textualists’ were wary of attributing meaning to historical processes or events, maintaining the gap between “aesthetics, on the one hand, grappling with the ‘permanent’, and sociology, on the other, investigating the ‘variable’” (Galan 1985), greater attention was placed on language as a process, unfolding in real time in a certain context—as opposed to a static, isolated structure. With striking similarities to New Criticism, Blaise Pascal in Pensées spoke of ‘recuperation’, to reconcile all contradictory passages in a text in order to elucidate its meaning (Easthope 1991, 21). Pierre Macherey disagreed with Pascal’s view of unity, saying such a text is organicist and vitalist. A text has an unconscious aspect found in “the lacunae and contradictions the text tries to smooth out” (ibid., 23). Macherey attributed a text with a conscious and unconscious component. Meaning is always plural: When we explain the work, instead of ascending to a hidden centre which is the source of life (the interpretative fallacy is organicist and vitalist), we perceive its actual decentredness. (Macherey 1978, 79)

The poststructuralist ideas that revise ‘structuralism’ take the reader into account, so that multiple meanings (interpretations) of a text are possible (Easthope 1988). It incorporates Lacan’s alignment of language with the unconscious. Reading is a process; a sentence unfolds in time and there is a dialectic between the reader and the text. Meaning is produced on reading, rather than being bound up in the text itself. This shift to poststructuralism was reflected by Julia Kristeva’s notion of subject positions in her ‘Revolution in poetic language’: “The realm of signification is always ‘a realm of positions … establishing the identification of the subject’” (Kristeva 1984, 43).

9  THE INTERPRETING ORGANISM 

199

Jacques Lacan introduced a bar between signifier and signified, in a similar way to Derrida’s detachment of the signifier from its Saussurean pairing (the connection between signifier/signified resembling the two sides of a sheet of paper according to Saussure) and the materiality of the signifier was emphasised. The signifier and its material processes are associated with the unconscious. The signifier, even in Saussure’s account, always had a residual materiality, namely, the sound or the mark on a page (Chandler 2007). With poststructuralism, this materiality took centre-­ stage and the signifier, therefore, assumed a greater autonomy. The process of signification was now envisioned as a chain of signifiers, signifier-signifier, which at certain points produced signified (concepts), or conscious thought. Lacan termed these meeting points of unconscious process and conscious meaning, point de capiton. Form and the unconscious were associated with the material signifier. Not surprisingly, such ideas were taken up by literary theorists inspired by Marxism: Saussurean linguistics had been revised so that now the material signifier was a determinant of thought itself.

Ideology and Realism Literary formalism differentiates literary from practical texts. In the former, the poetic function predominates as the focus is turned onto the ‘message’ (Chap. 6). A literary text is therefore ‘unmotivated’, while practical (or realist) texts are ‘motivated’. The latter are susceptible to ideology since they act as a transparent ‘window’, the vehicle by which extra-literary ideologically motivated causes are conveyed. Practical language, also termed standard language, becomes the property of interested classes: “A reflection of societal pressures, standard language is clearly subject to the contending claims of ‘aristocracy’ and ‘democracy’” (Galan 1985). The literary theorist, Mikhail Bakhtin differentiated the polyphonic text and its dialogism (Dostoyevsky’s work being the prime example), from the epic, which is a monologue and closed. The latter is susceptible to ideology, taking a similar position to that of the Althusserian poststructuralists that the realist text in literature is a subject of ideology (see below)—Catherine Belsey, for example, differentiated the interrogative from the realist text; the latter being a product of ideology. Theories developed in Prague insisted that there is no pure practical or literary text. A text has both: “Thus poetic language is intimately bound up with the standard language, but must always be distinguished from it”

200 

P. MCMAHON

(ibid.). They have different functions: standard (practical) language is communicative and poetic language, aesthetic. And consequently, one (practical language) aims to normalise expression and the other (poetic language) to be innovative, to realise “new possibilities” (ibid.). In this sense, Jakobson and the Prague school have continued the formalist opposition proposed by Lev Jakubinsky, a member of Opoyaz group in St Petersburg, that is, the division between literary and practical texts. But under the Prague revisions they are integrated into one system. Literariness, therefore, becomes relative (Chap. 6). Now a dialectic relation between actualisation and automatization (the terminology of the Prague theorists) underpins literary change. The former is “oblique” and resistant, while the latter facilitates understanding, “referring, unobstructed, to the objects under discussion” (ibid.). Poststructuralist theories developed in the 1970s and 80s went even further and denied transparency to language. Rather, some literary theorists claimed that apparent transparency in any text was an effect of ideology. This group of poststructuralists in Britain, influenced by Althusserian Marxism, turned their attention to the way realist texts are constructions that support neo-liberal (or capitalist) orders. Thus rather than the socialist realism that plagued formalists in early twentieth-century Russia, attention was turned to a more cryptic form of realism that upholds and endorses ideas of individuality, progress and valour in the face of hardship. In a Foucauldian sense, there is not much difference between the two ‘realisms’: they both entail the control of discourse by the powerful—the autocratic Communist party on the one hand, and neo-liberal concentrations of capitalist power, on the other. Nevertheless, these structuralist/ poststructuralist thinkers of 1970s and 80s Britain were committed to Marxist ideals, viewed as providing the most promising pathway to a new kind of socialist freedom. Realist Literature Modernist art was a reaction against the ideology that defined realism in literature. Analysing realist novels of the nineteenth century, Colin MacCabe gives an account of multiple narratives in the text which are drawn together by a hierarchy of voices and eventually the reader (MacCabe 1992). This closure was ideologically driven, a manipulation of form to make textual content appear ‘real’ and transparent, and the protagonists as autonomous and individually heroic. The modernist form attempted to

9  THE INTERPRETING ORGANISM 

201

cut through earlier ideology, presenting a perceiving, fragmented subject rather than the autonomous, triumphant one of Victorian realism. For those holding the reins of political power, literature should subscribe to the political, whether capitalist or socialist; a form of realism consequently enters literature. This ideological mechanism is based on the premise of cause and effect: external social discourse as the cause and literature as one of its effects. The endorsement of the autonomy of literature by formalists has the corollary that realism is viewed as actively produced, an ideological product. This crosses the political divide since, on the one hand, the realist novels of the nineteenth century were seen as an effect of capitalist ideology (e.g., in Colin MacCabe’s work, using Althusser’s ideas on ideology) and in Stalin’s regime, socialist realism emerged as the only acceptable form of literature (effectively putting an end to the lively formalist movement within Russia). While these two ‘realisms’ are different, with the capitalist form being expressed in more subtle terms, they have in common the assumption that an external ideology or belief in progress is prior to, and constitutive of, its literary outputs, which are its effects and serve to entrench the dominant political ideology. The poststructuralist analysis of realism could be useful in questioning the current (dominant) biological paradigm, that of bioengineering. The current reductionism that informs biological practices, envisages that increased knowledge of the (lifeless) molecules in living cells can be put to service to human ends. These ideas are particularly relevant to developing critiques of the neo-liberal ascendancy that characterises modern society, and the goals of biological applications, particularly manipulation of genomes. Rather than turning to Foucault in questioning discourses presented as ‘truths’, they examined the structure of realist texts, showing that what is presented as ‘real’ is just another construction. This view, of course, is not well regarded in the scientific community—science is after all a striving for a ‘truth’, a verifiable process that can be usefully applied to society. However, stances surrounding issues such as bioengineering (or genetic improvements) and other biological applications are vulnerable to all sorts of discourse—they are not straightforward, despite claims by the companies that promote them. The Althusserian-inspired poststructuralist movement aimed to expose claims to ‘truth’ in relation to such issues, as being just another discourse that valorises the status quo, that is, the protection of capitalist interests. Just as with formalism, this movement has been displaced by

202 

P. MCMAHON

postmodernist ideas, such as actor-network theory and other branches of poststructuralism. The Althusserian version developed in Britain was really a sort of structuralism, or ‘after-structuralism’, as Anthony Easthope quipped, and differs from more postmodernist, Marxist ideas such as those of Gilles Deleuze and Felix Guattari. These Deleuzian ideas reject a constitutive unconscious, replacing it with a rhizomatic model of interacting empiricities. The Althusserian poststructuralists, on the other hand, applied the imaginary/symbolic division of Lacanian psychoanalysis to texts to reveal how signification (the formal relations within a text) explains conscious effects, fantasies, and beliefs.

A Decentred Biology In literary theory, the shift from text to reader is reflected by the shift from structuralism to a linguistic poststructuralism (Easthope 1988). The reader, according to Julia Kristeva, assumes various subject positions offered by the text. The concept of the individual also comes under threat, as the body becomes the site of the inscription of competing discourses and practices. A non-unitary, dispersed and decentred subject emerges. The subject is reconstituted as a decentred assemblage of positions. Kristeva, a member of the well-known Tel Quel group (and literary magazine), which focused on contemporary structuralist and Maoist ideas in the 1960s and 70s, took up the ideas of the Russian literary theorist, Mikhail Bakhtin. While a critic of formalism, Bakhtin worked within its scope according to Kristeva, and tried to “transcend its limitations” (Oliver 1997, 64). To Bakhtin the word (or text) is not an isolated object, but is produced at the dynamic intersection of literary and extra-literary factors. The poetic word, therefore, is not a unitary entity in the text but is “polyvalent and multi-determined” (ibid., 65). Bakhtin could be considered a precursor of poststructuralism, and indeed was an inspiration for Kristeva, who herself is considered to be one of the major figures in the transition from structuralist to poststructuralist thought. Bakhtin, influenced by neo-Kantianism (especially the thought of Hermann Cohen) with its proposal that the grounding of existence is in the self/other opposition (Clark and Holquist 1984, 65), introduced the idea of dialogism. “Dialogism celebrates alterity” as the authority of the other is needed to define the self (ibid.). To Bakhtin, no individual can state a truth. Rather, it is only in the interaction with the other that meaningful statements can be made.

9  THE INTERPRETING ORGANISM 

203

Therefore, the polyphonic novel has a range of conflicting voices—only by engaging with other characters can a truth emerge. Like the poststructuralist subject, the separate, stand-alone organism, a basic concept to biology, is becoming decentred—the concept of the individual biological organism is under threat. The biological sciences are undergoing a paradigm shift as the ubiquity of symbiosis is becoming apparent (Simon et al. 2019). Symbiosis is a broad term as it includes various associations, including mutualism and parasitism. The endosymbiotic theory even proposes eukaryote cells (cells that include membranes, compartmentalisation and a nucleus) are the result of a symbiosis of prokaryotes—mitochondria, which have their own DNA, are proposed to be ancient (modified) bacteria. Now symbiosis is seen as essential to virtually all living communities. Lynn Margulis, who developed the theory of endosymbiosis, also introduced the term ‘holobiont’. Woodlands function on the symbiosis between trees and fungi (forming specialised root structures known as mycorrhizae). Corals are in symbiotic relations with photosynthetic zooxanthellae (or dinoflagellates, a group of unicellular plankton). Coral bleaching, a widespread problem in the tropics, is attributed to the breakdown of this symbiotic relation, a result of increased temperatures due to climate change. Concepts such as the holobiont, metagenome and microbiome present new models which deconstruct the self-contained individual and organism. Now it is apparent that most organisms cannot survive without symbionts. Organisms depend on other species in symbiotic associations, often between eukaryotes (complex organisms, including humans) and prokaryotes (bacteria and archaea). Ruminants require bacteria to digest cellulose. Termites can break down wood—again due to their bacterial symbionts. Humans also depend on communities of bacteria in the gut. Such associations “have profoundly challenged our traditional views of the physiology of nutrition, the performance of metabolic functions, and biological individuality itself, as they shifted the focus to the myriad of bacteria that enable organisms to perform these functions essential to their maintenance and survival” (Bognon-Kuss 2023, 196). The autonomy of the stand-alone organism is therefore being challenged in biology, as the identity and separate individuality of the organism are now under question. Cecelia Bognon-Kuss highlights the ways in which our views of the organism are changing. Under current conceptions of autonomy, a dialectical relation of the organism to its environment is postulated. By converting external materials, the organism establishes its

204 

P. MCMAHON

independent identity. While organisms are dependent on sources of external materials “it is precisely in this metabolic openness to the environment that organisms realize their autonomy and identity as they construct from them components that are their own as well as the organizational closure without which the system could not subsist as such” (ibid., 203). To Varela, Maturana and the relational biologists this defines metabolic closure (see Chap. 5). Metabolism has thereby drawn boundaries around the organism in this dialectical model—but now, Bognon-Kuss suggests, metabolism needs to be reconsidered in the light of entwined symbiotic relations between microorganisms and host organisms: “[T]he metabolism around which this conception of identity has been reinforced is now contributing to redraw its boundaries” (ibid., 205). Now the holobiont (the host and its symbionts) is considered by some to be the true ‘individual’ of biology. Process is now emphasised rather than the actual constituents of a community. This brings to mind the well-known textbook by Howard and Eugene Odum, Fundamentals of Ecology, which emphasised ecosystem function. The composition of an ecological community (its constituent species) may change, but its physiological functions (energy conversion and the processing of nutrient and waste materials) are conserved. Moreover, symbiotic associations challenge the dialectic model (in which the organism asserts both distinctiveness and persistence by engaging with the environment). This, according to Cecelia Bognon-Kuss, entails a separation of form and metabolic process—form is distinctive, while metabolism is persistent involving “the processes by which individual organisms maintain their identity over time” (ibid., 204). This could be considered a return to nineteenth-century concepts, including the Weismannian separation of germ line and somatic cells—the germ cells being definitive of form, while active body cells (the soma) metabolise. Such a separation needs to be deconstructed; this would involve a blurring of the form/matter distinction (ibid., 209). Bognon-Kuss proposes a shift from a self-centred view of the organism, to one based on ‘other’ and ‘change’ is needed. As “it has become clear that the isolated organism is incapable of functioning properly independently” (ibid., 207), new conceptions of the organism are needed—that introduce plurality, cooperation and otherness. This relational view repositions the organism as part of a network. However, a concept of otherness need not be the empirical-object centred networks of ANT, but could also be approached by linguistic models.

9  THE INTERPRETING ORGANISM 

205

Literary formalists attempted to integrate diachronicity with the synchronic system of Saussure, and these ideas were taken up by Lacan and some of the poststructuralist ideas he inspired.

References Barbieri, Marcello. 2013. Organic semiosis and Peircean semiosis. Biosemiotics 6 (2): 273–289. https://doi.org/10.1007/s12304-­012-­9161-­5. ———. 2019. Code biology, Peircean biosemiotics, and Rosen’s relational biology. Biological Theory 14 (1): 21–29. https://doi.org/10.1007/s13752-0180312-z Belsey, Catherine. 2002. Critical practice. 2nd ed. New York: Routledge. Bognon-Kuss, Cecelia. 2023. Metabolism in crisis? A new interplay between physiology and ecology. In Vitalism and its legacy in twentieth century life sciences and philosophy, ed. Christopher Donohue and Charles T.  Wolfe, 193–216. Cham: Springer. Chandler, D. 2007. Semiotics: The basics. 2nd. Taylor and Francis 2007 ed. London and New York: Routledge, Taylor and Francis Group. Original ed., 2002. Clark, Katerina, and Michael Holquist. 1984. Mikhail Bakhtin. Cambridge, MA: Belknap Press of Harvard University Press. Conley, Verena Andermatt. 1997. Ecopolitics: The environment in poststructuralist thought. London: Routledge. Dews, Peter. 1987. Logics of disintegration: Post-structuralist thought and the claims of critical theory. London, New York: Verso. Easthope, Antony. 1988. British post-structuralism since 1968. London: Routledge. ———. 1991. Literary into cultural studies. London, New York: Routledge. ———. 2002. Poetry as discourse, new accents. Routledge. Easthope, Antony, and Catherine Belsey. 2002. Privileging difference. Basingstoke: Palgrave. Fry, Paul H. 2012. Theory of literature, the open Yale course series. New Haven and London: Yale University Press. Galan, F.W. 1985. Historic structures. Vol. Paperback. Austin: University of Texas Press. Hoffmeyer, Jesper. 2008. Biosemiotics: An examination into the signs of life and the life of signs. Scranton: Scranton University Press. Kristeva, Julia. 1984. Revolution in poetic language. New  York: Columbia University Press. Kull, Kalevi. 1998. Semiotic ecology: Different natures in the semiosphere. Sign Systems Studies 26: 344–371. https://doi.org/10.12697/SSS.1998.26.15. Lash, Scott. 2002. Critique of information. Thousand Oaks, Calif: SAGE. Lewontin, R.C. 2000. The triple helix: Gene, organism, and environment. Cambridge: Harvard University Press.

206 

P. MCMAHON

MacCabe, Colin. 1992. Realism and the cinema: Notes on some popular Brechtian theses. In Modern literary theory: A reader, ed. Philip Rice and Patricia Waugh. London; New  York: E.  Arnold. (Distributed in the USA by Routledge, Chapman, and Hall, New York). Original edition, Screen 1974 Vol 15, 2. Macherey, P. 1978. A theory of literary production. Boston, London: Routledge and Kegan Paul. Oliver, Kelly, ed. 1997. The portable Kristeva. New York: Columbia University Press. Peirce, Charles Sanders. 1960. Collected papers Cambridge. Mass: Harvard University. Simon, Jean-Christophe, Julian R.  Marchesi, Christophe Mougel, and Marc-­ André Selosse. 2019. Host-microbiota interactions: From holobiont theory to analysis. Microbiome 7 (1): 5. https://doi.org/10.1186/s40168-­019-­0619-­4. Sullivan, Heather, and Bernhard Malkmus. 2016. The challenge of ecology to the humanities: An introduction. New German Critique 43: 1–20. https://doi. org/10.1215/0094033X-­3511835. Tønnessen, Morten. 2022. The evolutionary origin(s) of the umwelt. Biosemiotics 15: 1–5. https://doi.org/10.1007/s12304-­022-­09506-­7. Tønnessen, Morten, Timo Maran, and Alexei Sharov. 2018. Phenomenology and biosemiotics. Biosemiotics 11 (3): 323–330. https://doi.org/10.1007/ s12304-­018-­9345-­8. Weber, A. 2002. Feeling the signs: The origins of meaning in the biological philosophy of Susanne K. Langer and Hans Jonas. Sign Systems Studies 30 (1): 183. Wimsatt, W.K., and Monroe Beardsley. 1949. The affective fallacy. Sewanee Review 57 (1): 31–55.

CHAPTER 10

An Ecological Context

Questions Raised about Approaching Living Systems as Material for a Humanist ‘Cause’ The premises supported by current science and policy that indulge and support a bioengineering future are challenged by formalist and structuralist ideas. Literary formalism turns attention onto language itself, and its opaque, recalcitrant features. Its denial of language as an instrument, whether of unambiguous communication or some ideological cause, becomes a thorn in the side of discourses of power, which actively sideline it. Literary theory in general is under threat, as humanities departments are converted into a “province of the social sciences” (Khitrova 2019). Biological theory is also marginalised both by entrenched reductionism (in which organisation is overlooked while addressing the complex interactions of biocomponents) and by emerging postmodernist ideas that convert the relations that constitute life into networks of interacting objects. The human agent is uploaded to join other objects, as Bruno Latour put it. A new kind of reductionism is evident, one that turns to cybernetics and emergent properties. The marginalisation of alternative ideas by postmodern biology, with its explicit aim to develop bioengineering solutions, has something in common with the fate of literary formalism. A transparency is sought whereby biological systems can be interpreted in the final instance by physical processes and chemical components. The issue of biological organisation is © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3_10

207

208 

P. MCMAHON

overlooked or labelled as vitalist. Hence, biological theory investigating the constitutive relations of the organism is conflated with ideas of a vitalist ‘force’. In the hands of molecular reductionists such as Francis Crick, vitalist accusations were levelled at anyone who disagreed with a bioengineering future (Petersen 2023); this becomes an ideology in itself. The rise of systems biology is presented as way forward in elucidating the complex interactions and emergent properties that define living process, but in its way is equally reductionist and just as concerned with developing bioengineered organisms. Studies in the field of relational biology on organisational invariance reveal systemic properties, while ‘systems biology’ as currently practised concerns itself with a “vast list of components” and is forgetful of the role of the whole system (Cornish-Bowden and Cárdenas 2005). Ideas of form and structure developed by the Russian formalists and the Prague circle of linguists are striking in their refusal to endorse literature as being an effect of social and historical forces. Formalists refused to accept revolutionary principles as constituting a creative force in literature despite operating under the extreme political duress of Stalinism. Formalists turn around the notion of a transparent message, and concentrate on the medium itself; in other words, language is regarded as having its own being, other than just being a functional means of communication. and (parallel to this) the organism has a unique set of rules, or relations, that are not a mere outcome of function or need. A common feature of various expressions of formalism is that they spurn utopian visions, as well as an historicism underpinning an ideological cause: in the Soviet Union this was socialism and its flag bearer socialist realism, and in current neo-liberal societies it is the movement of capitalism into the realm of life itself (Cooper 2008). In the twentieth century, formalism and its incorporation into structuralist (systemic) models presented the most radical critique of such ideological movements. Now structuralist principles, in both literature and biology, have been actively marginalised. Structuralism is still implicit to the largely forgotten poststructuralist theory, developed in the 1970s and 80s, that incorporates principles of Althusser’s Marxist structuralism and Lacanian psychoanalysis. Meanwhile, alternative theories of life, such as relational biology, are also sidelined and have little influence on the new leviathan machine of bioengineering that embodies current biology. Formalist and structuralist views of relations in both the organism and its symbiotic relations, are seen to merely interfere with utopianesque, neo-liberal aspirations to control life itself, purportedly to benefit humanity.

10  AN ECOLOGICAL CONTEXT 

209

Modification: An External Cause Current evolutionary theory is at the heart of refashioning the organism as an ‘effect’, the outcome of a series of historically situated ‘causes’. The discourse underpinning the genetic manipulation of life rests on a certain premise endorsed in modern biology—this places the organism into a category shared by other phenomena in applied science as being in a causal relation with external events or circumstances. The proposed separation of germ line cells from the soma by August Weismann, as well as the archetype of Naturphilosophie, could be compared to the relation of author to text in expressive realism, favoured by the Romantic movement. The genetic system, therefore, becomes the ‘author’ of its passive product, the soma (or body). Within modern biology circles, the organism is considered a product—just as the products of a chemical reaction are derived from the predictable action of its reactants. Nonetheless, the chemical and physical processes in living cells are admittedly extremely complex (made apparent by the computing power required to account for the interactions between cellular components in modern systems biology). But the assumption remains that these processes are no different in the end to other chemical and physical processes (albeit with emergent properties). As discussed in Chap. 1, the logical empiricists of the Vienna circle held to this reductionist position, later taken up by Francis Crick and the ‘molecular revolution’ (Petersen 2023), claiming that opposing views are forms of vitalism. An emergentist model of systems is linked to Darwinian selection of mutated genes. These drive systems that account, in the end, for the organism and its ability to adapt to new environmental conditions, and therefore undergo change. While neo-Darwinism places its chips with mutations as the random, chance events behind adaptive changes, including the major structures and their relations (the ‘body plan’), structuralists disagree. Rather than attributing change to random alterations in the code of life, they focus instead on dynamic transformations within the system. Genetic manipulation and visions of a bioengineering future, such as synthetic biology, introduce an external cause—this time based on humanist goals. Formulated as the technologies of the future, they are presented as mere extensions of the proposal that the organism is an effect of physical and chemical processes. Resistance to such an interpretation can be found in formalist theories. Literary theories developed by formalists are united in ascribing meaning to formal relations within the text itself. Similarly, evolutionary developmental biology (evo-devo) places more

210 

P. MCMAHON

emphasis on the phenotype, and its internal relations, viewing genes as participants in the systems that constitute the organism. Ideas on literary evolution emerging from Russia and such evo-devo ideas, therefore, reject the alignment of internal change with an external cause. Evo-devo theorists connect phenotypic changes to unpredictable transformations that are generated internally and are subjected to natural selection. These ideas of unpredictable change are also consistent with formalist models of literary evolution. Both introduce the notion of ‘deformation’ and ‘accommodation’ as drivers of change. A further difference between neo-Darwinist and alternative evo-devo ideas relates to the notion of a progressive increase in complexity. Not only neo-Darwinist, but also some structuralist theories subscribe to an evolutionary momentum or a drive towards increased biological complexity. In the adaptationist account, this is attributed to singular, chance events that are selected additively and cumulatively. In orthogenetic accounts, such as those proposed by Lamarck and Haeckel, the drive is directed to a certain end (a legacy of the Great Chain of Being, a ladder-like progressive series from the simple to complex, with humans at the pinnacle). Alternative biological ideas proposed by Simon Conway-Morris and others, focus on convergence as an effect of external (unidentified) regularities or laws in nature. Convergence (also known as homoplasy) differs from homology in that there is no historical connection between very similar structures; the cephalopod (octopus) and vertebrate eye being one example (see Fig. 8.3). However, a convergent evolution that has only one outcome falls back into old theories of orthogenesis. Formalist concepts of historical change reject notions of a directed evolution and also question the notion of a tendency to increased complexity. Evidence is provided that atavistic forms return in literary genres, and similarly in phylogenetic series. In other words, simpler structures in early phylogenies can recur in more recent evolutionary branches. And similarly previous literary forms can make a come-back in more recent literature. Nevertheless, an evolutionary theory based on structural transformations (as opposed to mutations) might be restricted in the number of possible arrangements between constituent elements, and this might account for some instances of homoplasy (or convergence).

10  AN ECOLOGICAL CONTEXT 

211

Discourse of Modern Agriculture Generally, manipulative developments in agricultural and other technologies are viewed as the inevitable result of progress, and as the purpose of science. But by taking a critical stance, these unexamined assumptions can be questioned and viewed in the light of their context of social, religious and cultural beliefs. The technological ambition favoured currently in agriculture is the application of biotransformation technologies; genetic modification could improve domestic crops and animals by introducing promising candidate genes. These, it is proposed, will express their traits in the transformed organism and solve problems ranging from pest and drought susceptibility to low yields. Questions that might be asked by a history of science scholar are not whether these are ‘safe’ or whether they will ‘work’, although these are valid questions, but under which historical discourse have these technologies emerged as being a perfectly normal part of an advancing science? The usually unstated assumption is that science is an isolated, disinterested endeavour. Progress is made possible with increasing levels of knowledge and the advent of new techniques and equipment, such as the microscope in the seventeenth century. Discourse theorists, such as Michel Foucault, have questioned this assumption. Modern forms of knowledge, Foucault proposed, are structured by regularities common to different disciplines—thus each discipline (including biology) has common epistemological features with the others in the same historical period. Moreover, discourse is linked to power interests; in that sense, the biological sciences are not immune to the pressures of government, institutional and societal expectations. With this in mind, a question could be asked: Is modern agriculture, including its transformationist aspirations, not just another discourse that could be challenged by alternative approaches? Alternatives such as agroecology and its offshoots, permaculture and organic farming, could question assumptions of the dominant discourse of genetic manipulation. Such alternative ideas would interrogate the assumption of transformation being the ‘norm’, an inevitable outcome of progress. On the contrary, whilst the premise of biotransformation is that applying invasive technology is the best way to manage nature, agroecology is based on a level of reality in which interaction between ecological players (species, organisms, populations, holobionts) generate change—thus physiological changes in the system can be traced back to whole organism processes. An increased

212 

P. MCMAHON

nitrogen (N) content of soil, for example, may be the result of the fixation of atmospheric N into plant available forms of N by Rhizobium, a bacterium that associates symbiotically with roots of leguminous plants. Consequently, farmers plant legumes to improve soil fertility. The metabolic process is driven by the bacterium, but within the framework of symbiotic relations established with plant roots. The modern discourse of transformation is blind to this reality: it takes for granted that physiological processes can be taken out of their ecological context and applied in isolation to achieve desired results. In discourses promoting genetic manipulation, a perceived problem is isolated from its contextual setting and addressed with a single-hit technological solution. This raises another question that might be asked from an agroecological perspective: Why do some agrosystems have such severe pest problems in the first place? Is there another way to use resources available? The technologically driven discourse excludes such ideas. The only way forward, it is proposed, is to change the components of the system. Therefore, now the majority of cotton and maize crops in the United States are engineered for resistance to the herbicide, glyphosate. This has created resistance in weeds and resulted in increasing loads of chemicals in the environment. The removal of milkweed from the agricultural system, through herbicide applications, has been linked to a drastic decline in populations of the Monarch butterfly (Michigan University 2018). Other problems arise with the use of Bt crops, which expresses a toxin produced by a transgene inserted from a soil bacterium, Bacillus thuringiensis, that specifically targets Lepidopteran (moth) pests, including bollworm. As observed on Bt cotton in China, with the decline in bollworm moth pest populations, mirid pest populations increased—mirids were previously only a minor pest (ibid.). An agricultural system works according to ecological principles. Taking this into account, an integrated pest management (IPM) approach addresses the context of each pest, since pest populations are affected by the presence of other species in the agrosystem. The psychoanalyst Julia Kristeva observed a biological fatalism in the health sector based on the ‘pill’ instead of the ‘talk’ (Kristeva 1995). This is short-sighted as in therapy both a pharmaceutical and a therapeutic strategy are needed to work in concert. Agriculture is no different. Relying on technical solutions or isolated genetic transformations is not enough— we need to engage the whole system and treat organisms in the context of their ecological situation. Permaculture uses methods such as intercropping to establish barriers to the spread of pests or disease. Mixed cropping

10  AN ECOLOGICAL CONTEXT 

213

and rotation are other traditional methods used to deter pest/disease problems. In other words, an ecological approach can also be successful and is in the long term more sustainable, but it requires a new kind of organisation, an arrangement in production methods that is aligned as much as possible with natural processes. Agricultural problems can be addressed using an integrative, rather than reductionist approach. A critical difference from modern technological approaches is that these systems are aimed at the level of the whole organism and interactions between species, not underlying reduced forms of physiology. Integrated systems are also relevant to the urgent issue of climate change and greenhouse gas emissions since in most cases they sequester more carbon than crops grown in monoculture and require lower inputs of industrial inorganic fertilisers, further reducing emissions (McMahon and Keane 2023, 346).

Context of Environment Biotechnology removes (transformed) organisms from their context, although paradoxically the transgenes identified to engineer organisms are based on the responses of natural (untransformed) organisms to their conditions of existence. As recounted in Chap. 2, Georges Cuvier tied the organism to its conditions of existence. Under the modern synthesis of biology this was interpreted in terms of Mendelian genetics. An organism in a particular environment has a range of genes associated with fitness to survive that environment. That is the genes and associated phenotypic systems have been selected in response to particular environmental conditions. Biotechnology taps into this adaptationist thesis and attempts to identify major genes (or at least genes that account for a large portion of the variance in responses of the organism to its external conditions), and then transfer those genes to another, non-adapted species. Therefore, the biotechnological intervention transforms the organism from one in constant communication, responsive to its environmental conditions to one that expresses the added character despite its context. The character is now removed from its original context. The hope, of course, is that the expression of this character will replicate (or mimic) its original adaptive function. But the principle remains: the engineered organism is acontextual—it imposes a change onto the environment, rather than the two-way process between the internal and external that features in the original organism.

214 

P. MCMAHON

Take, for example, a drought-resistant species: resistance to drought involves many genes in different locations. However, some may be more important than others. If a significant quantitative trait locus (QTL) containing a gene can be identified, and that gene is transferred by engineering to a drought-sensitive species, the expectation is the transformed species will be more resistant to drought. But this resistance is no longer contextual—it is not an outcome of the response of the whole phenotype to drought. Rather, it depends on the uni-directionality of expression of the new gene in the transformed organism. The genetically manipulated organism is no longer contextual. As discussed below, employing traditional (quantitative) breeding practices can be used to enhance performance (such as resistance to pests, disease or drought) by the selection (in successive generations or by using molecular identification methods such as marker-assisted selection [MAS]) of a number of unrelated genes on different chromosomes. This type of resistance is broad and durable, as opposed to resistance relying on one or a few genes.

Quantitative Breeding: Contributions from Population Biology Should we be working with naturally evolved systems that are contextual or should we apply humanist principles, and impose our will onto the environment using modified species that can be productive regardless of their external situation? Our aspirations since the Renaissance speculations on manipulation by Francis Bacon has been to take the latter path. But alternative views have been expressed in biology and ecology, endorsing more holistic approaches to agriculture. The agricultural system is founded on ecological principles, and working against this may improve gains in the short-term, but in the long-term is a poor policy. Even the green revolution has been criticised on a number of fronts, not only by activists, but also by analysts working for the USDA (Ladejinsky 1977). Biotech simply takes these policies one step further: organisms must be transformed to meet human needs, removing them from the original ecological contexts in which they developed. This is made evident by the contrast between traditional (quantitative) breeding and direct (genetic) manipulation. Genetic engineering is not another kind of ‘breeding’ as its protagonists claim, since breeding in agriculture incorporates the principles of quantitative genetics.

10  AN ECOLOGICAL CONTEXT 

215

Traditional breeding, and its successor in modern agriculture quantitative breeding, aims to use heritable characters to improve crops and livestock—Darwin, who coined the term ‘artificial selection’, developed his ideas of natural selection from observations of the selection of promising characters in domestic breeds. Most phenotypic characters are associated with a range of different genes, in different locations of the genome. Traditional breeding (aimed to improve varieties for agricultural purposes) is not based on the genetic make-up per se, but on the phenotype and, importantly, on the phenotype in the context of its environment. Quantitative estimates of heritability take this into account: using statistical tools such as the analysis of variance and comparison of performance in different locations (therefore different environments), the variance attributed to genetic factors can be separated from environmentally influenced and background variation. Quantitative breeding is a selection process not of individual genes but of a number of genes and their related systems, via the phenotype—because it is the phenotype that is observed (e.g., photosynthetic rates, yields, resistance to pests and diseases). This leaves clear water between quantitative breeding and genetic manipulation (direct manipulation of the genome by inserting or knocking out genes) since breeding works with the whole organism while genetic engineering targets specific genes or sections of the genome (Fig. 10.1). In debates surrounding genetic transformation technologies, the point is raised that we manipulate organisms through breeding and have been doing so for millennia. ‘Molecular breeding’, the euphemistic description of genetic manipulation, is portrayed as being really the same in principle as traditional forms of breeding, but even more targeted and accurate. However, a critical difference is that modern genetic technologies seek to directly alter the organism at the level of the genome, while traditional methods always work with whole organisms. Biotechnologies have been around since ancient times but these always involved the whole organism: for example, fermentation technologies depended on bacterial or fungal species. Breeding crop plants involves crossing parents to produce an F1 generation, and backcrosses produce F2 progeny. This enables promising genes (e.g. from lines that show high yields and resistance to pests, diseases or drought) to be combined into one genome. All genes, their chromosomes and crossovers during meiosis, are involved at each stage of the breeding process. Following a number of such crosses, superior varieties with high yield and adaptations to the local environment can be produced. A report on genetically transformed crops prepared by Gurian-Sherman

216 

P. MCMAHON

A. Whole organism

Changed organism Quantitative genetics

Selection Traditional Breeding

Form

Context of environment Holistic regulation

Changed organism

B. Whole organism Genetic transformation Biological systems

Direct manipulation

Isolated genes, systems Decontextualised Reductionlist

Fig. 10.1  Traditional selection and breeding manipulate organisms for agriculture, fermentation in wine-making and other activities, working with whole organisms, which act as ‘gatekeepers’ to changes, while modern technologies employ direct manipulation, particularly genetic transformation technologies, bypassing the determining relations of the whole organism

(2009) concluded that yield increases in broad-acre crops were attributable to traditional crossing methods and not transformed genes. The main ‘success’ of transgenic crops was due to their herbicide resistance, enabling broad scale application of herbicides to control weeds. This has inevitably led to an increased chemical load in the environment with ecological consequences (see above). Molecular methods also play an important part in modern agricultural breeding methods, but in a role that does not involve direct manipulation. Modern molecular tools are becoming important in breeding. For example, molecular markers (often collected in a ‘library’) can be used to estimate the genomic location of promising traits (quantitative trait loci), saving the time and money needed to incorporate characteristics by breeding successive generations. Quantitative breeding, therefore, increasingly makes use of molecular technologies but retains the centrality of the whole organism in the process of breeding.

10  AN ECOLOGICAL CONTEXT 

217

In traditional breeding the organism (its formal relations) acts as a ‘gatekeeper’. Changes are incorporated within the context of the whole. As illustrated in Fig. 10.1, both methods (traditional breeding and genetic transformation) arrive at a ‘changed organism’ but in traditional breeding, characteristics determined by a range of genes are incorporated through the complicated process of meiosis by crossing successive generations (or more rapidly by using molecular tools in marker-assisted selection), while genetic transformation inserts genes expressing isolated functions directly into the host DNA. In traditional methods, the organism is involved in each step of the process, while genetic transformation results in a changed organism only at the end. Traditional breeding has generally been ‘quantitative’. Quantitative breeding, now advanced by the molecular determination of quantitative trait loci (QTLs), is holistic in that phenotypic characters are selected to establish promising crosses (e.g., between varieties showing pest or disease resistance and high yield). The promising characters in each parent are determined not by a single gene, but by the combined effect of a number of genes (or more accurately systems of gene expression). Quantitative breeding attempts to increase the frequency of these promising genes—in this sense it is anti-reductionist, while molecular ‘breeding’ follows a more reductionist model. An additional feature is that so long as quantitative breeding is conducted in the context of its environment, the performance of the crop or domestic animal takes environmental parameters into account. These can be even more important than the genetic component: killing off pollinators by the use of pesticides could be counterproductive, for example, as we cannot expect to improve yields in insect-pollinated crops if the pollinators are impacted (McMahon and Keane 2023, 216). Interplanting genetically diverse lines of rice helps to control important diseases, such as rice blast disease in China (Zhu et al. 2000). The structural heterogeneity of the system plays a role in disease control as well, demonstrated by Chinese research where tall and short rice varieties were interplanted to reduce the impact of rice blast. Further considerations are climate, micro-­ climate, soil conditions and the list goes on. On the contrary, genetic manipulation is often decontextualised, both from the organism as a whole and various environmental parameters. In addition to the myth that genetic manipulation is the modern, improved and more accurate version of breeding), another myth is that traditional breeding is anti-technological, inspired by an unrealistic romantic vision of organic communities in the past. A distinction therefore is put

218 

P. MCMAHON

forward that what separates modern agriculture from traditional breeding is the application of molecular biology methods (including genetic transformation). Molecular methods and genetic manipulation are conflated into one seemingly uncontentious unit. This is misleading. For one it suggests that genetic manipulation automatically follows in the wake of molecular technologies, as if it is an unquestionable consequence of improved technology. For another, it presents an image of modern (advanced) technological methods as replacing obsolete, backward, traditional practices. The narrative is that only by turning agriculture over to genetic transformation technologies, can we ensure food security in the future. However, the real distinction (as shown in Fig. 10.1) is between technologies of direct manipulation (genetic transformation) and technologies that depend on the formal relations of the organism to incorporate new characters, such as resistance to crop diseases or drought. Molecular methods such as marker-assisted selection (MAS) contribute enormously to quantitative breeding methods. In each case, whether the traditional farmer selecting his or her best seeds, or a modern breeder identifying suitable parents for crossing using molecular methods, the desired change has to pass the barrier of the whole organism. Changes are incorporated by the process of meiosis. In contrast, genetic transformation bypasses this ‘gatekeeping’ role to directly implement the change. Genetic transformation depends on the function of one or a few genes, while quantitative breeding may incorporate a range of genes (Keane 2012). Characters linked to a range of genes are more robust and durable than those dependent on single genes. Quantitative breeding generally occurs in the context of the external environment: it seeks the additive effect of a number of unrelated genes to improve performance in a particular context. Genetic transformation in contrast is conducted independently of any environmental context. Behind this is the unsupported assumption of nature ultimately succumbing to the scalpel of science, while basic questions such as Schrodinger’s ‘What is life?’ remain unanswered. Treating natural forms as ‘other’ and not as objects to assimilate into humanist aspirations could lead, in itself, to more sustainable solutions in agriculture or other spheres of activity.

10  AN ECOLOGICAL CONTEXT 

219

References Cooper, Melinda E. 2008. Life as Surplus: Biotechnology and Capitalism in the Neoliberal Era. Seattle, WA: University of Washington Press. Cornish-Bowden, Athel, and M.  Cárdenas. 2005. Systems Biology May Work When We Learn to Understand the Parts in Terms of the Whole. Biochemical Society Transactions 33: 516–519. https://doi.org/10.1042/BST0330516. Gurian-Sherman, Doug. 2009. Failure to Yield. Union of Concerned Scientists. Keane, P.J. 2012. Horizontal or Generalized Resistance to Pathogens in Plants InTech-Horizontal_or_generalized_resistance_to_pathogens_in_plants.pdf. In Plant Pathology, edited by Christian Joseph R. Cumagun. IntechOpen. Khitrova, Daria. 2019. Introduction to Permanent Evolution. In Permanent Evolution: Selected Essays on Literature, Theory and Film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 1–24. Boston: Academic Studies Press. Kristeva, Julia. 1995. New Maladies of the Soul. New  York: Columbia University Press. Ladejinsky, Wolf. 1977. 52. Agriculture-Some Broader Considerations. In Agrarian Reform as Unfinished Business: The Selected Papers of Wolf Ladejinsky, ed. Louis J. Walinsky, 462. Oxford University Press. Original edition, Economic Situation and Prospects of India—April 21, 1970. McMahon, Peter, and Philip Keane. 2023. Chocolate and sustainable cocoa farming. Newcastle, UK: Cambridge Scholars Publishing. Petersen, Eric L. 2023. A ‘Fourth Wave’ of Vitalism in the Mid-20th Century? In Vitalism and Its Legacy in Twentieth Century Life Sciences and Philosophy, ed. Christopher Donohue and Charles T. Wolfe, 173–192. Switzerland: Springer. University, Michigan State. 2018. Superweeds, Secondary Pests and Lack of Biodiversity are Frequent GMO Concerns. Futures. Zhu Y., Chen H., Fan J., Wang Y., Li Y., Chen J., et al. 2000. Genetic diversity and disease control in rice. Nature 406 (6797): 718–722. https://doi.org/ 10.1038/35021046.

References

Adami, C., C. Ofria, and T.C. Collier. 2000. Evolution of Biological Complexity. Proceedings of the National Academy of Sciences of the United States of America 97 (9): 4463–4468. https://doi.org/10.1073/pnas.97.9.4463. Appel, Toby. 1987. The Cuvier-Geoffroy Debate. Oxford University Press. Bahm, Archie J. 1983. Five Systems Concepts of Society. Behavioral Science 28 (3): 204–218. https://doi.org/10.1002/bs.3830280304. Barbieri, Marcello. 2013. Organic Semiosis and Peircean Semiosis. Biosemiotics 6 (2): 273–289. https://doi.org/10.1007/s12304-­012-­9161-­5. ———. 2019. Code biology, Peircean biosemiotics, and Rosen’s relational biology. Biological Theory 14 (1): 21–29. https://doi.org/10.1007/s13752-0180312-z Belsey, Catherine. 2002. Critical Practice. 2nd ed. New York: Routledge. Bergson, H. 1911. Creative Evolution. London: MacMillan and co. Bognon-Kuss, Cecelia. 2023. Metabolism in Crisis? A New Interplay between Physiology and Ecology. In Vitalism and Its Legacy in Twentieth Century Life Sciences and Philosophy, ed. Christopher Donohue and Charles T.  Wolfe, 193–216. Switzerland: Springer. Bolker, Jessica A. 2000. Modularity in Development and Why It Matters to Evo-­ Devo. American Zoologist 40 (5): 770–776. https://doi.org/10.1668/0003-­ 1569(2000)040[0770:MIDAWI]2.0.CO;2. Boucher, Sandy C. 2015. Functionalism and Structuralism as Philosophical Stances: van Fraassen Meets the Philosophy of Biology. Biology and Philosophy 30 (3): 383–403. https://doi.org/10.1007/s10539-­014-­9453-­z.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3

221

222 

REFERENCES

Bradford, Richard. 1994. Roman Jakobson: Life, Language, Art. London and New York: Routledge. Braendle, Christian, and Thomas Flatt. 2006. A Role for Genetic Accommodation in Evolution? BioEssays 28 (9): 868–873. https://doi.org/10.1002/ bies.20456. Brown, Edward J. 1974. The Formalist Contribution. The Russian Review (Wiley) 33 (3): 243–258. Capra, Fritjof. 1997. The Web of Life: A New Synthesis of Mind and Matter. London: Flamingo. Cardenas, M.L., J.C.  Letelier, C.  Gutierrez, A.  Cornish-Bowden, and J.  Soto-­ Andrade. 2010. Closure to Efficient Causation, Computability and Artificial Life. Journal of Theoretical Biology 263 (1): 79–92. Carrol, Sean B., Jennifer K. Grenier, and Scott D. Weatherbee. 2001. From DNA to Diversity: Molecular Genetics and the Evolution of Animal Design. Oxford: Blackwell Science Ltd. Cassirer, Ernst. 1945. Structuralism in Modern Linguistics. WORD 1: 99–120. Chandler, D. 2007. Semiotics: The Basics. 2nd ed. Taylor and Francis 2007 ed. London and New  York: Routledge, Taylor and Francis Group. Original edition, 2002. Chapman, Anne. 2005. Genetic engineering: The unnatural argument. Techne 9, 2: Winter 2005 pp 81–93. Chirimuuta, M. 2023. The Critical Difference Between Holism and Vitalism in Cassirer’s Philosophy of Science. In Vitalism and Its Legacy in Twentieth Century Life Sciences and Philosophy, ed. Christopher Donohue and Charles T. Wolfe, 85–106. Switzerland: Springer. Clark, Katerina, and Michael Holquist. 1984. Mikhail Bakhtin. Cambridge, MA: Belknap Press of Harvard University Press. Colebrook, Claire. 2002. Gilles Deleuze. New York: Routledge. Conley, Verena Andermatt. 1997. Ecopolitics: The Environment in Poststructuralist Thought. London: Routledge. Conway Morris, S. 2003. Life’s Solution: Inevitable Humans in a Lonely Universe. Cambridge: Cambridge University Press. Cooper, Melinda E. 2008. Life as Surplus: Biotechnology and Capitalism in the Neoliberal Era. Seattle, WA: University of Washington Press. Cornish-Bowden, A., and M.L.  Cárdenas. 2005. Systems Biology May Work When We Learn to Understand the Parts in Terms of the Whole. Biochemical Society Transactions 33: 516–519. https://doi.org/10.1042/BST0330516. ———. 2007. Organizational Invariance in (M,R)-Systems. Chemistry & Biodiversity 4 (10): 2396–2406. ———. 2008. Self-organization at the Origin of Life. Journal of Theoretical Biology 252 (3): 411–418. https://doi.org/10.1016/j.jtbi.2007.07.035.

 REFERENCES 

223

———. 2020. Contrasting Theories of Life: Historical Context, Current Theories. In Search of an Ideal Theory. Biosystems 188: 104063. https://doi. org/10.1016/j.biosystems.2019.104063. ———. 2022. The Essence of Life Revisited: How Theories Can Shed Light on It. Theory in Biosciences = Theorie in den Biowissenschaften 141 (2): 105–123. https://doi.org/10.1007/s12064-­021-­00342-­w. Cornish-Bowden, Athel, María Luz Cárdenas, Juan-Carlos Letelier, Jorge Soto-­ Andrade, and Flavio Guíñez Abarzúa. 2004. Understanding the Parts in Terms of the Whole. Biology of the Cell 96 (9): 713–717. https://doi.org/10.1016/j. biolcel.2004.06.006. Cornish-Bowden, A., M.L. Cardenas, J.C. Letelier, and J. Soto-Andrade. 2007. Beyond Reductionism: Metabolic Circularity as a Guiding Vision for a Real Biology of Systems. Proteomics 7 (6): 839–845. Crispo, Erika. 2007. The Baldwin Effect and Genetic Assimilation: Revisiting Two Mechanisms of Evolutionary Change Mediated By Phenotypic Plasticity. Evolution 61 (11): 2469–2479. https://doi.org/10.1111/j.1558-­5646.2007 .00203.x. Cuccato, G., G. Della Gatta, and D. di Bernardo. 2009. Systems and Synthetic Biology: Tackling Genetic Networks and Complex Diseases. Heredity 102 (6): 527–532. https://doi.org/10.1038/hdy.2009.18. Culler, Jonathan. 1976. Saussure. London: Fontana. Darwin, Charles. 1859. On the Origin of Species. 1st ed. Harvard University Press. Dawkins, R. 1991. The Blind Watchmaker. 2006 edition. London: Penguin. Denton, M.J. 2013. Types: A Persistent Structuralist Challenge to Darwinian Pan-­ selectionism. BIO-Complexity 3: 1–18. https://doi.org/10.5048/BIO-­C. 2013.3. Dews, Peter. 1987. Logics of Disintegration: Post-Structuralist Thought and the Claims of Critical Theory. London and New York: Verso. Driesch, Hans, 1892 [1964]. “Entwicklungsmechanische Studien. I. Der Werth der beiden ersten Furchungszellen in der Echinodermentwicklung. Experimentelle Erzeugen von Theil- und Doppelbildung”, Zeitschrift für wissenschafliche Zoologie, 53: 160–178. Abridged translation as “The Potency of the First Two Cleavage Cells in Echinoderm Development. Experimental Production of Partial and Double Formations.” L. Mezger, M. Hamburger, V. Hamburger, and T.S. Hall (trans.), in Willier and Oppenheimer 1964: 38–50. Eagleton, Terry. 1983. Literary Theory: An Introduction. Oxford: Basil Blackwell Publisher. Easthope, Antony. 1988. British Post-structuralism since 1968. London: Routledge. ———. 1991. Literary into Cultural Studies. London and New York: Routledge. ———. 2002. Poetry as Discourse, New Accents. Routledge. Easthope, Antony, and Catherine Belsey. 2002. Privileging Difference. Basingstoke, [UK]: Palgrave.

224 

REFERENCES

Ekstig, B. 2015. Complexity, Natural Selection and the Evolution of Life and Humans. Foundations of Science 20 (2): 175–187. https://doi.org/10.1007/ s10699-­014-­9358-­y. Eldredge, N. and Gould, S.J. 1972. Punctuated equilibria: an alternative to phyletic gradualism. In Schopf, T.J.M. (ed.). Models in Paleobiology. San Francisco, CA: Freeman Cooper. pp. 82–115. Erlich, Victor. 1969. Russian Formalism: History-Doctrine. The Hague: Mouton. Evans, B.R., P.  Kotsakiozi, A.L.  Costa-da-Silva, R.S.  Ioshino, L.  Garziera, M.C.  Pedrosa, A.  Malavasi, J.F.  Virginio, M.L.  Capurro, and J.R.  Powell. 2019. Transgenic Aedes aegypti Mosquitoes Transfer Genes into a Natural Population. Scientific Reports 9 (1): 13047. https://doi.org/10.1038/ s41598-­019-­49660-­6. Fagan, Melinda Bonnie, and Jane Maienschein. 2022. Theories of Biological Development. In The Stanford Encyclopedia of Philosophy (Summer 2022 Edition), ed. Edward N. Zalta. Stanford: Metaphysics Research Lab, Stanford University. Ferry, Luc. 1995. The New Ecological Order. Translated by Carol Volk. Chicago: University of Chicago Press. Foucault, Michel. 1970. The Order of Things: An Archaeology of the Human Sciences. New York: Pantheon Books. Frazzetta, T.H. 1975. Complex Adaptations in Evolving Populations. Sunderland, MA: Sinauer. Fry, Paul H. 2012. Theory of Literature, The Open Yale Course Series. New Haven and London: Yale University Press. Galan, F.W. 1979. Literary System and Systemic Change: The Prague School Theory of Literary History, 1928–48. Proceedings of the Modern Language Association (PMLA) 94 (2): 275–285. ———. 1985. Historic Structures. Vol. Paperback. Austin: University of Texas Press. Gancedo, C., and C.L. Flores. 2008. Moonlighting Proteins in Yeasts. Microbiology and Molecular Biology Reviews 72 (1): 197–210. https://doi.org/10.1128/ mmbr.00036-­07. Gatherer, D. 2010. So what do we really mean when we say that systems biology is holistic? BMC Systems Biology 4 (1): 1–12. https://doi.org/10.1186/ 1752-0509-4-22. Gerber, Sylvain, and Melanie J.  Hopkins. 2011. Mosaic Heterochrony and Evolutionary Modularity: The Trilobite Genus Zacanthopsis as a Case Study. Evolution 65 (11): 3241–3252. https://doi.org/10.1111/j.1558-­5646.2011 .01363.x. Gilbert, Scott F. 2003. Evo-Devo, Devo-Evo, and Devgen-Popgen. Biology and Philosophy 18: 347–352. Goodchild P. 1996. Deleuze and Guattari: An Introduction to the Politics of Desire. USA: SAGE Publications.

 REFERENCES 

225

Goodwin, B.C. 1990. Structuralism in Biology. Science Progress (1933–), 47: 227–244. ———. 1994. How the Leopard Changed Its Spots: The Evolution of Complexity. London: Charles Scribner’s Sons. Goodwin, B.C., and Christian Brière. 1991. Generic Dynamics of Morphogenesis. https://doi.org/10.1142/9789812819598_0007. Gould, S.J. 1977. Ontogeny and phylogeny. Cambridge, MA and London, UK: Belknap Press (Harvard UP). ———. 1994. Hen’s Teeth and Horse’s Toes. Paperback edition. USA: Norton. ———. 2002. The Structure of Evolutionary Theory. Cambridge, MA: Belknap Press of Harvard University Press. Goyal, R.K., M.A.  Schmidt, and M.F.  Hynes. 2021. Molecular Biology in the Improvement of Biological Nitrogen Fixation by Rhizobia and Extending the Scope to Cereals. Microorganisms 9 (1). https://doi.org/10.3390/micro organisms9010125. Greene, John C. 1981. Science, Ideology and World View. University of California Press. Grene, Marjorie, and David J. Depew. 2004. The Philosophy of Biology: An Episodic History. New York: Cambridge University Press. Griffiths, Anthony J.F., Jeffrey H. Miller, David T. Suzuki, Richard C. Lewontin, and William M.  Gelbart. 2000. An Introduction to Genetic Analysis. 7th ed. W.H. Freeman & Company. Gruber, S., C.P. Kubicek, and V. Seidl-Seiboth. 2011. Differential Regulation of Orthologous Chitinase Genes in Mycoparasitic Trichoderma Species. Applied and Environmental Microbiology 77 (20): 7217–7226. https://doi.org/10. 1128/aem.06027-­11. Guardian, The. 1999. Top Researchers Back Suspended Lab Whistleblower. The Guardian, 12 February 1999. Gurian-Sherman, Doug 2009. Failure to yield. Union of Concerned Scientists Report. Cambridge MA: UCS Publications. Habermas, Jurgen. 2003. The Future of Human Nature. Polity Press. Hall, Brian K. 2003. Evo-Devo: Evolutionary Developmental Mechanisms. The International Journal of Developmental Biology 47 (7–8): 491–495. https:// doi.org/10.1387/ijdb.14756324. Haraway, Donna. 1997. Modest_Witness@Second_Millennium.FemaleMan_Meets_ OncoMouse: Feminism and Technoscience. New York: Routledge. Haraway, Donna Jeanne. 1991. Simians, Cyborgs, and Women: The Reinvention of Nature. London: Free Association. ———. 2003. The Haraway Reader. New York: Routledge. Heidegger, Martin. 1995. The Fundamental Concepts of Metaphysics. Translated by William McNeill and Nicholas Walker. Bloomington: Indiana University Press.

226 

REFERENCES

Ho, Mae Wan. 2000. Genetic Engineering: Dream or Nightmare? Continuum Intl Pub Group; Revised, Updated 2000. Hoffmeyer, Jesper. 2008. Biosemiotics: An Examination into the Signs of Life and the Life of Signs. Scranton: Scranton University Press. Huntley, Herbert E. 1970. The Divine Proportion. New York: Dover Publications. Huxley, Julian. 1942. Evolution: The Modern Synthesis. Harper and Brothers. Jablonka, Eva, and Marion Lamb. 1995. Epigenetic Inheritance and Evolution: The Lamarckian Dimension. Oxford and New York: Oxford University Press. ———. 2005. Evolution in Four Dimensions. Genetic, Epigenetic, Behavioral, and Symbolic Variation in the History of Life. Cambridge, MA and London: MIT Press. Jackson, Wes. 1984. A New Threat to Agriculture and a Window of Opportunity. The Ecologist 14 (3): 119–124. Jakobson, Roman. 1971a. Two Aspects of Language and Two Types of Aphasic Disturbances. In Roman Jakobson Selected Writings Volume II Word and Language. Berlin and New York: De Gruyter Mouton. ———. 1971b. Word and Language. In Roman Jakobson Selected Writings Volume II Word and Language. Berlin and New York: De Gruyter Mouton. Original edition, 1929. ———. 1987. Linguistics and Poetics. In Language in Literature, ed. Krystyna Pomorska and Stephen Rudy. Cambridge, MA: The Belknap Press of Harvard University Press. Jameson, F. 1991. Postmodernism, or, the cultural logic of late capitalism. USA: Duke University Press. Jeffery, Constance J. 2003. Moonlighting Proteins: Old Proteins Learning New Tricks. Trends in Genetics 19 (8): 415–417. https://doi.org/10.1016/S0168-­ 9525(03)00167-­7. Jonas, Hans. 1966. The Phenomenon of Life: Toward a Philosophical Biology. New York: Harper and Row. ———. 1974. Philosophical Essays from Ancient Creed to Technological Man. Prentice Hall. Karsenti, Eric. 2008. Self-organization in Cell Biology: A Brief History. Nature Reviews Molecular Cell Biology 9 (3): 255–262. https://doi.org/10.1038/ nrm2357. Keane, P.J. 2012. Horizontal or Generalized Resistance to Pathogens in Plants in Tech-Horizontal_or_generalized_resistance_to_pathogens_in_plants.pdf. In Plant Pathology, ed. Christian Joseph R. Cumagun. IntechOpen. Keller, Evelyn Fox. 2000. Century of the Gene. Cambridge, MA: Harvard University Press. Khitrova, Daria. 2019. Introduction to Permanent Evolution. In Permanent Evolution: Selected Essays on Literature, Theory and Film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 1–24. Boston: Academic Studies Press.

 REFERENCES 

227

Kristeva, Julia. 1984. Revolution in Poetic Language. New  York: Columbia University Press. ———. 1995. New Maladies of the Soul. New York: Columbia University Press. Kull, Kalevi. 1998. Semiotic Ecology: Different Natures in the Semiosphere. Sign Systems Studies 26: 344–371. https://doi.org/10.12697/SSS.1998.26.15. Kull, Kalevi, Claus Emmeche, and Jesper Hoffmeyer. 2011. Why Biosemiotics? An Introduction to Our View on the Biology of Life Itself. In Towards a Semiotic Biology 1–21 Singapore: World Scientific DOI: 10.1142/P771. Ladejinsky, Wolf. 1977. 52. Agriculture-Some Broader Considerations. In Agrarian Reform as Unfinished Business: The Selected Papers of Wolf Ladejinsky, ed. Louis J. Walinsky, 462. Oxford University Press. Original edition, Economic Situation and Prospects of India—April 21, 1970. Larsen, Ellen. 2003. Genes, Cell Behavior and the Evolution of Form. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 119–132. Cambridge, MA: MIT Press. Lash, Scott. 2002. Critique of Information. Thousand Oaks, CA: SAGE. ———. 2006. Life (Vitalism). Theory, Culture and Society 23 (2–3): 323–329. https://doi.org/10.1177/0263276406062697. Latour, Bruno. 2005. Reassembling the Social: An Introduction to Actor-Network-­ Theory. Oxford: Oxford University Press. Le Guyader, Hervé. 1998 [2004]. Etienne Geoffroy Saint-Hilaire, 1772–1844: Un Naturaliste Visionnaire, (Un Savant, Une Époque), Paris: Belin. Translated as Étienne Geoffroy Saint-Hilaire, 1772–1844: A Visionary Naturalist, 2004. Chicago: University of Chicago Press. Letelier, Juan Carlos, Gonzalo Marı́n, and Jorge Mpodozis. 2003. Autopoietic and (M,R) Systems. Journal of Theoretical Biology 222 (2): 261–272. https:// doi.org/10.1016/S0022-­5193(03)00034-­1. Letelier, Juan-Carlos, María Luz Cárdenas, and Athel Cornish-Bowden. 2011. From L’Homme Machine to Metabolic Closure: Steps Towards Understanding Life. Journal of Theoretical Biology 286: 100–113. https://doi.org/10.1016/j. jtbi.2011.06.033. Levin, M. 2012. Morphogenetic Fields in Embryogenesis, Regeneration, and Cancer: Non-Local Control of Complex Patterning. Biosystems 109 (3): 243–261. https://doi.org/10.1016/j.biosystems.2012.04.005. Lewontin, R.C. 2000. The Triple Helix: Gene, Organism, and Environment. Cambridge: Harvard University Press. MacCabe, Colin. 1992. Realism and the Cinema: Notes on Some Popular Brechtian Theses. In Modern Literary Theory: A Reader, ed. Philip Rice and Patricia Waugh. London and New York: E. Arnold (Distributed in the USA by Routledge, Chapman, and Hall, New  York). Original edition, Screen 1974 Vol. 15, 2.

228 

REFERENCES

Macherey, P. 1978. A Theory of Literary Production. Boston and London: Routledge and Kegan Paul. Matthews, Eric. 1996. Twentieth-century French Philosophy. New  York: Oxford University Press. Maynard-Smith, John. 1966. Sympatric Speciation. The American Naturalist 110 (916): 637–650. https://doi.org/10.1086/282457. McMahon, Peter, and Philip Keane. 2023. Chocolate and Sustainable Cocoa Farming. Newcastle, UK: Cambridge Scholars Publishing. Merchant, C. 1980. The death of nature: Women, ecology and the scientific revolution. USA: Harper and Row. Milner, Andrew. 1991. Contemporary Cultural Theory: An Introduction. North Sydney: Allen & Unwin. Moir, Cat. 2023. What Is Living and What Is Dead in Political Vitalism? In Vitalism and Its Legacy in Twentieth Century Life Sciences and Philosophy, ed. Christopher Donohue and Charles T. Wolfe, 239–262. Switzerland: Springer. Monod, J. 1971. Chance and Necessity: an Essay on the Natural Philosophy of Modern Biology. New York: Knopf. Mukarovosky, Jan. 1932. Poetry and Standard Language (Basnictvi a Spisovny jazyk). In Spisnova Cestina a Jazykova Kultura, ed. Havranek, Bohuslav and Weingart, Milos, 123–256. Translated by FW Galan. Praha: Melantrich. Muller, Gerd. 2003. Homology: The Evolution of Morphological Organisation. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 51–70. Cambridge, MA: MIT Press. Muller, Gerd, and Stuart Newman, eds. 2003. Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. Edited by Gerd Muller, Gunter Wagner and Werner Callebaut. The Vienna Series in Theoretical Biology. Cambridge, MA: MIT Press. Nanjundiah, Vidyanand. 2003. Phenotypic Plasticity and Evolution by Genetic Assimilation. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 245–264. Cambridge, MA: MIT Press. Oliver, Kelly, ed. 1997. The Portable Kristeva. New York: Columbia University Press. Peirce, Charles Sanders. 1960. Collected papers Cambridge. Mass: Harvard University. Petersen, Eric L. 2023. A ‘Fourth Wave’ of Vitalism in the Mid-20th Century? In Vitalism and Its Legacy in Twentieth Century Life Sciences and Philosophy, ed. Christopher Donohue and Charles T. Wolfe, 173–192. Switzerland: Springer. Piaget, J. 1971. Structuralism. London: Routledge & Kegan Paul. Pigliucci, M., C.J. Murren, and C.D. Schlichting. 2006. Phenotypic Plasticity and Evolution by Genetic Assimilation. The Journal of Experimental Biology 209 (Pt 12): 2362–2367. https://doi.org/10.1242/jeb.02070.

 REFERENCES 

229

Pomorska, Krystyna. 1988. Jakobson Dialogues. Paperback edition (1980, 1983, 1988) ed. Cambridge, MA: MIT Press. Price, C.S.C. 1995. Structurally Unsound. Evolution, 49: 1298-1302. https:// doi.org/10.1111/j.1558-5646.1995.tb04463.x. Prindle, D. F. 2009. Stephen Jay Gould and the politics of evolution. USA: Prometheus Books. Protevi, John. 2015. Canguilhem’s ‘Comparative Physiology’: An Eco-Bio-Social Multiplicity. Symposium: Canadian Journal of Continental Philosophy/Revue Canadienne de Philosophie Continentale: 19(2): 57–71. Raff, R.A. 1996. The Shape of Life. Chicago: Chicago University Press. Rieppel, Olivier. 1990. Structuralism, Functionalism, and the Four Aristotelian Causes. Journal of the History of Biology 23 (2): 291–320. Rose, M. R., and Oakley, T. H. 2007. The new biology: Beyond the modern synthesis. Biology Direct 2: 30. https://doi.org/10.1186/1745-6150-2-30. Rosen, R. 1991. Life Itself: A Comprehensive Inquiry into the Nature, Origin, and Fabrication of Life. New York: Columbia University Press. ———. 2000. Essays on Life Itself. Edited by Timothy F.  H. Allen and David W.  Roberts, Complexity in Ecological Systems Series. New  York: Columbia University Press. Rosengrant, Sandra. 1980. The Theoretical Criticism of Jurij Tynjanov. Comparative Literature 32 (4): 355–389. Ruse, M. 1973. The philosophy of biology. London: Hutchinson and Co (Ltd). Ruse, M. 1996. Monad to Man: The Concept of Progress in Evolutionary Biology. Cambridge, MA: Harvard University Press. Russell, Edward Stuart. 1916. Form and Function. London: John Murray. Saussure, Ferdinand de. 1959. Course in General Linguistics. Translated by Wade Baskin. New York: Philosophical Library. Shepard, P. 1967. Man in the Landscape: A Historic View of the Esthetics of Nature. Athens, GA: University of Georgia Press. Shklovsky, V. 1992. Art as Technique. In Modern Literary Theory: A Reader, ed. Philip Rice and Patricia Waugh. London and New York: E. Arnold (Distributed in the USA by Routledge, Chapman, and Hall, New York). Simon, Jean-Christophe, Julian R. Marchesi, Christophe Mougel, and Marc-­André Selosse. 2019. Host-Microbiota Interactions: From Holobiont Theory to Analysis. Microbiome 7 (1): 5. https://doi.org/10.1186/s40168-­019-­0619-­4. Slijper, E.J. 1942. Biological Anatomical Investigations on the Bipedal Gait and Upright Posture in Mammals, with Special Reference to a Little Goat, Born without Forelegs. I. Proceedings of the Koninklijke Nederlandse Akademie Wetenschappen 45: 288–295. Stjernfelt, Frederik. 2011. Simple Animals and Complex Biology: Von Uexküll’s Two-Fold Influence on Cassirer’s Philosophy. Synthese 179 (1): 169–186. https://doi.org/10.1007/s11229-­009-­9634-­5.

230 

REFERENCES

Sullivan, Heather, and Bernhard Malkmus. 2016. The Challenge of Ecology to the Humanities: An Introduction. New German Critique 43: 1–20. https:// doi.org/10.1215/0094033X-­3511835. Suzuki, Y., and H.F.  Nijhout. 2006. Evolution of a Polyphenism by Genetic Accommodation. Science 311 (5761): 650–652. https://doi.org/10.1126/ science.1118888. Theißen, Günter. 2006. The Proper Place of Hopeful Monsters in Evolutionary Biology. Theory in Biosciences 124: 349–369. ———. 2009. Saltational Evolution: Hopeful Monsters are Here to Stay. Theory in Biosciences 128 (1): 43–51. https://doi.org/10.1007/s12064-­009-­0058-­z. Thompson, D. A. W. 1917. On growth and form. Cambridge: Cambridge University Press. Tipton, K.F., M.I. O’Sullivan, G.P. Davey, and J. O’Sullivan. 2003. It Can Be a Complicated Life Being an Enzyme. Biochemical Society Transactions 31 (Pt 3): 711–715. https://doi.org/10.1042/bst0310711. Todorov, T. 1978. L’Heritage formaliste. Jakobson, Cahiers Cistre (Laussanne) pp. 47–51. Tønnessen, Morten. 2022. The Evolutionary Origin(s) of the Umwelt. Biosemiotics 15: 1–5. https://doi.org/10.1007/s12304-­022-­09506-­7. Tønnessen, Morten, Timo Maran, and Alexei Sharov. 2018. Phenomenology and Biosemiotics. Biosemiotics 11 (3): 323–330. https://doi.org/10.1007/s12304-­ 018-­9345-­8. Turing, Alan M. 1952. The Chemical Basis of Morphogenesis. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 237 (641): 37–72. Twitty, V.C. 1932. Influence of the Eye on the Growth of its Associated Structures, Studied by Means of Heteroplastic Transplantations. Journal of Experimental Zoology 61: 333–375. Tynianov, Jurii. 2019. On Literary Evolution. In Permanent Evolution: Selected Essays on Literature, Theory and Film—Yuri Tynianov, ed. Ainsley Morse and Philip Redko, 267–282. Boston: Academic Studies Press. Tynianov, Jurii, and Roman Jakobson. 1972. Problems in the Study of Language and Literature. In The Structuralists: From Marx to Levi-Strauss, ed. Richard T. De George, 80–83. Garden City, NY: Anchor Books. University, Michigan State. 2018. Superweeds, Secondary Pests and Lack of Biodiversity are Frequent GMO Concerns. Futures. Venditti, Chris, Andrew Meade, and Mark Pagel. 2009. Phylogenies Reveal New Interpretation of Speciation and the Red Queen. Nature 463: 349–352. https://doi.org/10.1038/nature08630. Wagner, G. P. 1988. The influence of variation and of developmental constraints on the rate of multivariate phenotypic evolution. Journal of Evolutionary Biology 1 (1): 45–66.

 References 

231

Wagner, Gunter. 1999. Homologues, Natural Kinds, and the Evolution of Modularity. American Zoologist 36. https://doi.org/10.1093/icb/36.1.36. Waugh, Linda R. 1980. The Poetic Function in the Theory of Roman Jakobson. Poetics Today 2 (1a): 57–82. Weber, A. 2002. Feeling the signs: The origins of meaning in the biological philosophy of Susanne K. Langer and Hans Jonas. Sign Systems Studies 30 (1): 183. Webster, G. 1989. Structuralism and Darwinism: Concepts for the Study of Form. In Dynamic Structures in Biology, ed. Brian C. Goodwin, Atuhiro Sibatani, and Gerry Webster, 1–15. Edinburgh: Edinburgh University Press. Webster, G., and B.C.  Goodwin. 1982. The Origin of Species: A Structuralist Approach. Journal of Social and Biological Structures 5 (1): 15–47. https:// doi.org/10.1016/S0140-­1750(82)91390-­2. ———. 1996. Form and Transformation: Generative and Relational Principles in Biology. Cambridge UK and New York: Cambridge University Press. Wessel, M. 1971. Fertilizer requirements of cacao (Theobroma cacao L.) in southwestern Nigeria. Amsterdam: Koninklijk Institut voor de Tropen. West-Eberhard, M.J. 2003. Developmental plasticity and evolution. Oxford University Press. ———. 2005a. Developmental Plasticity and the Origin of Species Differences. Proceedings of the National Academy of Sciences of the United States of America 102 (Suppl 1): 6543–6549. https://doi.org/10.1073/pnas.0501844102. ———. 2005b. Phenotypic Accommodation: Adaptive Innovation Due to Developmental Plasticity. Journal of Experimental Zoology: Part B, Molecular and Developmental Evolution 304 (6): 610–618. https://doi.org/10.1002/ jez.b.21071. Willmer, Pat. 2003. Convergence and Homoplasy. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology, ed. Gerd Muller, Gunter Wagner, and Werner Callebaut, 33–49. Cambridge, MA: MIT Press. Wimsatt, W.K. Jr. and Monroe C. Beardsley. 1946. The Intentional Fallacy, in W. K. Wimsatt, Jr., The Verbal Icon (Lexington: University of Kentucky Press, 1967), p. 4. ———. 1949. The Affective Fallacy. Sewanee Review 57 (1): 31–55. Wolfe, Charles. 2011. From Substantival to Functional Vitalism and Beyond: Animas, Organisms, and Attitudes. Eidos 14: 212–235. Zambelli, Federico, Giulio Pavesi, Carmela Gissi, David S. Horner, and Graziano Pesole. 2010. Assessment of Orthologous Splicing Isoforms in Human and Mouse Orthologous Genes. BMC Genomics 11 (1): 534. https://doi.org/1 0.1186/1471-­2164-­11-­534. Zhu Y., Chen H., Fan J., Wang Y., Li Y., Chen J., et al. 2000. Genetic diversity and disease control in rice. Nature 406 (6797): 718–722. https://doi.org/ 10.1038/35021046.

Index

A Actor-Network Theory (ANT), 8 Actualisation and automatization, 127 as a dialectical response, 127 Adaptation, 29, 30, 35, 40, 43, 46 and adaptivity, 135–137 Adaptationism, 29, 32 Affective fallacy The Intentional Fallacy, 187 and the reader, 187, 188 Agroecology mixed cropping, 212 and permaculture, 211, 212 Alternative splicing, 158, 159 Althusser, Louis, 196, 201 Aphasia and linguistic development in children, 132 and metaphor and metonym poles, 124, 131 two aphasic types, 131 Appel, Toby, 30, 31, 36, 38, 39, 42–45

Aristotle and Cuvier’s biology, 33, 41 and species, 33 Artificial life (AL), 9, 10, 12 Automatization and actualization, 127 deformation, 128 B Bacon, Francis, 214 and Baconian, 30, 75 Bakhtin, Mikhail, 98 and dialogism, 151, 199 Baldwin effect, 174 Barbieri, Marcello, 189, 194, 195 Bateson, William, 166 Bauplan, 50, 65 Beardsley, Monroe, 78, 79 Belsey, Catherine, 79, 81, 86 and interrogative texts, 79 Bergson, Henri, 15, 19, 23, 24 and elan vital, 19, 22, 23 Bertalanffy, Ludwig von, 59 and general systems theory, 59

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 P. McMahon, Structuralism and Form in Literature and Biology, https://doi.org/10.1007/978-3-031-47739-3

233

234 

INDEX

Biosemiotics, 188–196 Bognon-Kuss, Cecelia, 203, 204 Bonnet, Charles, 33 Breeding in agriculture, 7 and quantitative methods, 5 Brenner, Sydney, 81 Brooks, Cleanth, 95 Bt crops, 212 Bacillus thuringiensis, 212 Buffon, Comte de, 32, 33 Bukharin, Nicolai, 100 C Camera eye, 183, 184 convergence (homoplasy), 183–185 Canalisation, 175, 176 Canguilhelm, Georges, 52 Capra, Fritjof, 71 Carbon sequestration, 213 Cardenas, Maria, 59 Carnap, Rudolf, 22 Cassirer, Ernst, 12–14, 21–23 and criticism of vitalism, 21–23 and presentation to New York Linguistics Circle, 12 Catalysts and infinite regress, 107–109 as intermediates in reactions, 108 Central Dogma, 57, 60, 64 Chain of being (scala naturae), 36, 38 Chapman, Anne, 1, 2, 8 Chomsky, Noam, 196 Circumnutation, 133, 134 Comparative anatomy, 34, 37, 42 Complexity evolution of, 143–148, 153 Conditions of existence, 29, 30, 34, 36–39, 42, 43, 46 and Cuvier’s Correlation of parts and Subordination of characters, 39

Constructive principle, 128 and deformation, 128 Convergent evolution (homoplasy), 170, 171, 183–185 Conway Morris, Simon, 24 convergent evolution, 24 Cooper, Melinda, 1, 3 Coral bleaching, 203 Cornish-Bowden, Athel, 59 Crick, Francis, 9–11, 22, 24 and criticism of vitalism, 22, 24 Cuvier, Georges, 20, 21, 23 and comparative anatomy, 21 and conditions of existence, 15 modern episteme, 34, 37–39 1830 debate, 29 Cuvier/Geoffroy Debate (1830), 29–46 D Darwin, Charles, 13, 15 and early evolutionary ideas, 15 and Erasmus, 72 Dawkins, Richard, 71, 81 de Vries, Hugo, 166 Defamiliarisation, 96 Deleuze, Gilles, 24, 25 Depew, David, 20 Derrida, Jacques, 188, 196, 197, 199 Dews, Peter, 119 Driesch, Hans, 12, 19, 20, 22, 23 and entelechy, 12, 19, 20, 22, 23 and epigenesis, 20 E Easthope, Antony, 188, 196–198, 202 Efficient causation, 103, 105, 108, 109 and metabolic closure, 103, 105 Eikhenbaum, Boris, 93, 96 Eldredge, Niles, 168

 INDEX 

Eliot, T.S., 80 Embranchement (phylum), 33, 36, 39, 44 Epigenesis, 20 as anti-theologian, 143 Epigenetic inheritance, 84 Episteme classical, 32, 33, 36 and Foucault, 30, 34, 38 modern, 34 Erlich, Victor, 100 Evolution combinatorial, 146, 158 evo-devo proposal of radical phenotypic change, 147, 151 immanent evolution, 139–147 of literature, 139–142, 148 and neo-Darwinism, 64, 72, 85, 115–116, 171, 209 Evolutionary developmental biology (evo-devo), 42, 46 Exaptation, 83 Exons, 158, 159 and introns, 158 F Fisher, Ronald A., 167 Florkin, Marcel, 194 Foucault, Michel, 30–34, 36, 38–40, 43 and The Order of Things, 31, 34 Fry, Paul, 94 Functionalist biology, 69 adaptationism, 69 Futurist poets, 99 G Galan, F.W., 117, 119, 124, 127, 130 Gene therapy, 6, 7 somatic vs. germ line, 6

235

Genetic accommodation, 165, 171, 174–179 and genetic assimilation, 165, 176–178 Genetic manipulation and bioengineering, 1–25 quantitative breeding, 5, 7 and somatic and germline cells, 2, 6, 7, 24 Genetic recombination as ephemeral, 165 and meiosis, 81 quantitative breeding, 75 Goethe, Johann von, 13, 21, 23 Goldschmidt, Richard, 168, 180–182 and ‘hopeful monsters,’ 168, 180, 181 Goodwin, Brian, 17, 20 Gould, Stephen Jay, 29 Green Revolution, 3 Greene, John, 33, 36–37, 41 Grene, Marjorie, 20, 21 H Habermas, Jurgen, 2, 7 Haeckel, Ernst, 24 and embryology, 36 and hylozoism, 24 Haraway, Donna, 25 Heidegger, Martin, 23 Helmsjlev, Louis, 195, 196 Heritability, 215 Heterochrony, 164, 169, 182 Homeobox (Hox) genes, 14 Homology and Darwin’s theory of descent, 30, 32, 33, 43, 61–63, 183 and DNA sequences, 6 and embryology, 20 pentadactyl limb, 35, 40, 77, 83, 153 orthologous genes, 157 Humboldt, Wilhelm von, 13

236 

INDEX

Huntley, Herbert E., 18 Hylozoism, 23, 24 panvitalism, 23, 24 I Infinite regress, 107–109 and enzymes, 107–108 Integrated pest management (IPM), 212 Intentional fallacy, 78 J Jablonka, Eva, 7 Jackson, Wes, 1–4, 6 Jakobson, Roman, 88, 89 and aphasia studies, 88 and metaphor and metonym poles, 89 Jakubinsky, Lav, 93 Jameson, Fredric, 73 Jonas, Hans, 104 K Kant, Immanuel, 11, 21, 23 and structural unity, 11 Kauffman, Stuart, 147 Keller, Evelyn Fox, 57 Khitrova, Daria, 79 Khlebinkov, Velimir, 99 and futurism, 99, 100 Kielmeyer, Carl, 42, 179 law of compensation, 42 Kristeva, Julia, 198, 202 Kull, Kalevi, 192 L Lacan, Jacques, 198, 199, 205 and British post-structuralist theory, 196 and psychoanalysis, 196, 197, 202

Ladejinsky, Wolf, 3 Lamarck, Jean-Baptiste, 20, 23, 24 and evolution of complexity, 23 Lamb, Marion, 7 Langer, Susanne, 135 Langue and parole, 16 and revision by Prague linguistics circle, 12 Larsen, Ellen, 160 Lash, Scott, 8, 24, 25 Latour, Bruno, 8 Levi-Strauss, Claude, 117, 121–123 Lewontin, Roger, 73, 74 Linnaeus, Carl, 33, 39 Literariness contrast to practical language, 93, 98, 99 Literary devices, 93, 98, 99 Literary formalism and authorial intention, 77, 78, 88 and autonomy, 77, 89 and literary evolution, 82, 96, 116, 124, 128, 129, 142, 145, 148, 167, 174, 180, 182, 184, 210 New Criticism, 77–80, 85 Prague structuralism, 79 Russian formalists, 77, 79, 80, 82 Lovelock, James, 193 M M/R systems, 101–103, 107 MacCabe, Colin, 200, 201 Macherey, Pierre, 198 Macroevolution, 165, 167, 169, 179, 181, 182 and modularity, 167, 181, 182 Margulis, Lynn, 203 Marker-assisted selection (MAS), 7 May, Todd, 136 Mayakovsky, Vladimir, 99, 100

 INDEX 

Maynard-Smith, John, 172, 173 Mayr, Ernst, 10 Meiosis, 215, 217, 218 Mendel, Gregor, 166 Metaphor and metonym, 124, 131, 133 and paradigmatic and syntagmatic axes, 119, 120, 124, 126, 132, 135 Modernist movement, 14–17 Modern synthesis of Darwinian biology and genetics, 72–75 Modular systems (modularity) and plasticity, 165, 169, 172 and recurrence, 165, 171, 177 watchmaker parable, 170 Morgan, Thomas Hunt, 166 Morphogenesis and bioelectric fields, 153 and chemical gradients, 153, 154 Morphology, 31, 40, 42 Moscow Linguistics Circle, 96 Mosquitoes genetic manipulation for disease control, 5 and Umwelt, 21 and US Environmental Protection Authority, 5 Muller, Gerd, 147, 156, 157 Multi-functionality and genes, 105–107 in language according to the Prague school, 99 and moonlighting proteins, 108, 110 Mutationism, 166–168, 181 opposition to gradualism, 167 N Nagel, Ernest, 22 Natural selection

237

as cumulative following genetic mutations, 71 and genetic accommodation, 57, 157, 165, 171, 174 and mimicry, 70, 71 Naturphilosophie, 64 Neo-Darwinism, 64, 72, 85, 115–116, 171, 209 New Criticism and affective fallacy, 79, 187–188 and intentional fallacy, 17 Newman, Stuart, 147, 157 Nitrogen fixation and genetic manipulation, 4 and legumes, 4 Non-computability, 109–110 and simulable models of artificial life, 109 O Oken, Lorenz, 65 Ontogeny, 42 and phylogeny, 84 Opoyaz, 96 Organisational invariance, 108 Orthologous genes (orthologs), 158–160 Overproduction and microtubules, 133, 134 in phenotype, 133–135 and searching behaviour, 134 Owen, Richard, 52 P Paley, William, 30 Pascal, Blaise, 198 Patents, 3 and genetic manipulation, 3 Peirce, Charles, 193–195 and triadic sign, 193, 195 Petersen, Erik, 9, 11, 21, 22

238 

INDEX

Phenomenology, 188, 191, 197 Phenotypic accommodation (compensation), 174–179 Phenotypic plasticity and developmental constraints, 116 and morphic types, 163 and overproduction, 134–137 Piaget, Jean, 80 Pleiotropy, 110 Poetic function, 115–137 and six functions, 126, 132 Polymorphism and heterochrony, 164 and population biology, 164 Post-structuralism critique of realist texts, 199–201 and transparency, 200 Prague Linguistics Circle, 117, 123–130 Preformation, 20 and Church authority, 20 Prindle, David, 74 Protevi, John, 136 Punctuated equilibria, 168–169 Pusztai, Arpad, 69 Q Quantitative trait loci (QTLs), 214, 216 R Rashevsky, Nicolas, 101, 102 Realism (literary), 201 and ideology, 199–202 Recapitulation and Haeckel, 24, 145, 156, 165, 210 and von Baer, 85 and von Baer, archetypal form, 85

Recurrence, 163, 165, 168, 170, 171, 177, 183–185 Relational biology and complex systems, 101 critique of systems biology, 105–107 and metabolic closure, 103, 105–110 Riedl, Rupert, 72 Rosen, Robert, 11, 12 Russell, Edward S., 13 Russian formalists, 77, 79, 80, 82 and defamiliarization, 96, 124, 127, 142, 151 and New Critics, 78–80 S Saint-Hilaire, Etienne Geoffroy, 14 and Cuvier-Geoffroy debate, 49 and Unity of Composition, 49, 64, 87, 149 Saint-Hilaire, Isidore Geoffroy, 29, 30, 35 morphological structures, 40–42 Saltation, 42 Saltationism, 166, 168, 171, 174, 180–183 Saussure, Ferdinand de, 15, 16 Sebeok, Thomas, 194 Self-organisation, 55 Shklovsky, Viktor, 96–99 Structuralism definition by Roman Jakobson, 117, 119, 130 high structuralism, 121 and Saussure’s Course in Linguistics, 115, 124 Structuralist biology, 13, 14, 18 Symbiosis, 203 holobiont, 192, 203, 204 Symbologists, 96 Synchronic language, 117, 119, 125

 INDEX 

opposition to diachronic or historical models, 117 Synthetic biology, 6, 9, 10 Systems biology, 58–60 T Teleology, 145 orthogenesis vs. non-­ directionality, 146–147 Thompson, D’Arcy, 10, 12, 15, 18, 19, 22 Transformism Geoffroy and saltationism, 36, 42, 45, 46 Lamarck’s theory, 36 Trotsky, Leon, 100 critique of formalism, 100 Turing, Alan, 153, 154 Tynianov, Yuri, 79, 80, 88, 89 coexistence of the present and the past, 146 and Jakobson, 124, 143 and literary evolution, 82, 96, 116, 124, 128, 129, 142, 145, 148, 167, 174, 180, 182, 184, 210 U Uexkull, Jakob von, 21, 22 Umwelt, 21 and biosemiotics, 105, 188–194

Unity of composition, 31, 33, 35, 36 and Geoffroy’s Principle of connection and Balance of parts, 39 V Vienna circle, 21, 22 Vitalism, 1–25 and organicism, 18–23 Voloshinov, Valentin, 100 von Baer, Karl and Darwin, 84–85 embryology, 85 and idealism, 85 von Baer’s rule, 172–174 W Waddington, Conrad, 14, 22 Wagner, Gunther, 158 Wallace, Alfred, 52 Webster, Gerry, 20 Weismann, August, 55, 64 Weismann barrier, 55, 64 and neo-Weismannian reduction, 80 West-Eberhard, Mary Jane, 14 Willmer, Pat, 157 Wilson, E.O., 73 Wimsatt, William, 78 Wolfe, Charles, 18, 20, 21

239